1:
2: \documentclass[11pt]{article}
3:
4: \input psfig.sty
5: \psfull
6:
7: \topmargin 0.25truein
8: \oddsidemargin 0.0truein
9: \evensidemargin 0.0truein
10: \textheight 8.5truein
11: \textwidth 6.5truein
12: \footskip 0.6truein
13: \headheight 0.0truein
14: \headsep 0.0truein
15: \parskip 0.3cm
16:
17: \begin{document}
18:
19: \newtheorem{lemma}{Lemma}
20: \newtheorem{theorem}{Theorem}
21: \newtheorem{corollary}{Corollary}
22: \newtheorem{definition}{Definition}
23: \newtheorem{example}{Example}
24: \newtheorem{proposition}{Proposition}
25: \newtheorem{condition}{Condition}
26:
27: \newcommand{\be}{\begin{equation}}
28: \newcommand{\ee}{\end{equation}}
29: \newcommand{\bea}{\begin{eqnarray}}
30: \newcommand{\eea}{\end{eqnarray}}
31: \newcommand{\beaa}{\begin{eqnarray*}}
32: \newcommand{\eeaa}{\end{eqnarray*}}
33: \newcommand{\ben}{\begin{enumerate}}
34: \newcommand{\een}{\end{enumerate}}
35: \newcommand{\bi}{\begin{itemize}}
36: \newcommand{\ei}{\end{itemize}}
37:
38: \newcommand{\lip}{\langle}
39: \newcommand{\lan}{\langle}
40: \newcommand{\rip}{\rangle}
41: \newcommand{\ran}{\rangle}
42: \newcommand{\uu}{\underline}
43: \newcommand{\oo}{\overline}
44: \newcommand{\til}{\tilde}
45:
46: \newcommand{\La}{\Lambda}
47: \newcommand{\la}{\lambda}
48: \newcommand{\eps}{\epsilon}
49: \newcommand{\vph}{\varphi}
50: \newcommand{\al}{\alpha}
51: \newcommand{\bet}{\beta}
52: \newcommand{\gam}{\gamma}
53: \newcommand{\Gam}{\Gamma}
54: \newcommand{\kap}{\kappa}
55: \newcommand{\Del}{\Delta}
56: \newcommand{\Th}{\Theta}
57: \newcommand{\s}{\sigma}
58: \newcommand{\sig}{\sigma}
59: \newcommand{\Sig}{\Sigma}
60: \newcommand{\del}{\delta}
61: \newcommand{\om}{\omega}
62: \newcommand{\Om}{\Omega}
63: \newcommand{\X}{{\bf \Xi}}
64:
65: \newcommand{\integers}{Z\!\!\!Z}
66: \newcommand{\Z}{Z\!\!\!Z}
67: \newcommand{\N}{{\rm I\!N}}
68: \newcommand{\rationals}{{\rm I\!Q}}
69: \newcommand{\reals}{{\rm I\!R}}
70: \newcommand{\R}{{\rm I\!R}}
71: \newcommand{\RJ}{{{\rm I\!R}^J}}
72: \newcommand{\RJP}{{{\rm I\!R}^J_+}}
73: \newcommand{\naturals}{{\rm I\!N}}
74:
75: \newcommand{\calA}{{\cal A}}
76: \newcommand{\calB}{{\cal B}}
77: \newcommand{\calC}{{\cal C}}
78: \newcommand{\calD}{{\cal D}}
79: \newcommand{\calF}{{\cal F}}
80: \newcommand{\calG}{{\cal G}}
81: \newcommand{\calH}{{\cal H}}
82: \newcommand{\calJ}{{\cal J}}
83: \newcommand{\calL}{{\cal L}}
84: \newcommand{\calM}{{\cal M}}
85: \newcommand{\calP}{{\cal P}}
86: \newcommand{\calS}{{\cal S}}
87: \newcommand{\calT}{{\cal T}}
88: \newcommand{\calU}{{\cal U}}
89: \newcommand{\calV}{{\cal V}}
90: \newcommand{\calX}{{\cal X}}
91: \newcommand{\calY}{{\cal Y}}
92:
93: \newcommand{\proof}{\noindent {\bf Proof:\ }}
94: \newcommand{\proofOf}[1]{\noindent {\bf Proof of #1:\ }}
95: \newcommand{\remark}{\noindent {\bf Remark:\ }}
96: \newcommand{\remarks}{\noindent {\bf Remarks:\ }}
97: \newcommand{\note}{\noindent {\bf Note:\ }}
98: \newcommand{\esssup}{{\rm ess}\sup}
99: \newcommand{\essinf}{{\rm ess}\inf}
100: \newcommand{\pl}{\partial}
101: \newcommand{\noi}{\noindent}
102: \newcommand{\goto}{\to}
103: \newcommand{\ink}{\rule{.5\baselineskip}{.55\baselineskip}}
104: \newcommand{\qed}{\rule{.5\baselineskip}{.55\baselineskip}}
105:
106: \def\ve{\varepsilon}
107: \def\vr{\varrho}
108:
109: \newcommand{\diag}{{\rm diag}}
110: \newcommand{\trace}{{\rm trace}}
111: \newcommand{\tr}{{\rm tr}}
112: \newcommand{\w}{\wedge}
113: \newcommand{\dint}{\int\!\!\!\int}
114: \newcommand{\lt}{\left}
115: \newcommand{\rt}{\right}
116: \newcommand{\dist}{{\rm dist}}
117:
118: \newcommand{\policy}{{u}}
119: \def\OBdry{\partial_o}
120: \def\CBdry{\partial_c}
121: \newcommand{\Sfrac}[2]{{{#1}\slash {#2}}}
122:
123: \title{
124: An escape time criterion for queueing networks:
125: Asymptotic risk-sensitive control via differential games%
126: \thanks{This
127: research was supported in part by the United States---Israel Binational
128: Science Foundation (BSF 1999179)}}
129: \author{Rami Atar\footnote{Electrical Engineering,
130: Technion---Israel Institute of Technology,
131: Haifa 32000, Israel. Research of this author also supported
132: in part by the fund for promotion of research at the Technion.},
133: Paul Dupuis\footnote{Lefschetz Center for Dynamical Systems,
134: Brown University,
135: Division of Applied Mathematics,
136: Providence, R.I.\ 02912.
137: Research of this author also supported in part by
138: the National
139: Science Foundation (NSF-DMS-0072004, NSF-ECS-9979250) and the Army Research Office
140: (DAAD19-99-1-0223).}
141: \ and Adam Shwartz\footnote{Electrical Engineering,
142: Technion---Israel Institute of Technology,
143: Haifa 32000, Israel. Research of this author also supported
144: in part by INTAS grant 265, and in part by the fund for promotion
145: of research at the Technion.}\\[.2in]
146: }
147:
148: %
149: \date{February 2, 2003}
150:
151: \maketitle
152:
153: \begin{abstract}
154: We consider the problem of risk-sensitive control of a stochastic network.
155: In controlling such a network,
156: an escape time criterion can be useful if one wishes
157: to regulate the occurrence of large buffers and buffer overflow.
158: In this paper a risk-sensitive escape time criterion is formulated,
159: which in comparison to the ordinary escape time criteria penalizes exits which occur on short time intervals more heavily.
160: The properties of the risk-sensitive problem are studied in the large buffer limit,
161: and related to the value of a deterministic differential game with constrained dynamics.
162: We prove that the game has value, and that the value
163: is the (viscosity) solution of a PDE.
164: For a simple network, the value is computed,
165: demonstrating the applicability of the approach.
166: \end{abstract}
167:
168: \section{Introduction}
169:
170: In this paper we consider a problem of risk-sensitive control
171: (or rare event control) for queueing networks.
172: The network includes
173: servers that can offer service to two or more
174: classes of customers, and a choice must be made
175: regarding which classes to offer service at each time.
176: We study a stochastic control
177: problem in which this choice is regarded as the control, and where
178: the cost is a risk-sensitive version of the time to
179: escape a bounded set. Hence, fixing $c>0$,
180: and denoting by $\s$ the time when the queue-lengths process
181: first exits a given domain, we consider $E_xe^{-c\s}$ as a criterion
182: to be minimized.
183: Such a criterion penalizes short exit times more heavily
184: than ordinary escape time criteria
185: (such as $E_x\s$, a criterion to be maximized).
186: There are at least two motivations for the use of such criteria when designing policies for the control of a network.
187: The first is that in many communication networks system performance is measured in terms of rare event probabilities (e.g., probabilities of data loss or excessive delay).
188: The second motivation follows from the connection between risk-sensitive controls and robust controls.
189: Indeed, as discussed in \cite{dupjampet},
190: the optimization of a single fixed stochastic network with respect to a risk-sensitive cost criteria automatically produces controls with specific and predictable robust properties.
191: In particular,
192: these controls give good performance for a family of perturbed network models (where the perturbation is around the design model and the size of the perturbation is measured by relative entropy), and with respect to a corresponding ordinary (i.e., not risk-sensitive) cost.
193:
194: In many problems,
195: one considers the limit of the risk-sensitive problem as a scaling parameter of the system converges,
196: in the hope that the limit model is more tractable.
197: We follow the same approach here,
198: and show that the normalized costs in the risk-sensitive problems converge to the value function of a differential game with constraints.
199: As is well known,
200: the convergence analysis is closely related to the large deviation properties of the sequence of controlled processes.
201: An interesting feature in the setting of stochastic networks is that the asymptotic analysis of a sequence of {\it controlled} networks is in many ways simpler than the analogous asymptotic analysis of a sequence of {\it uncontrolled} networks.
202: For example,
203: if one were to fix a particular state feedback service policy at each station,
204: then the calculation of the large deviation asymptotics is very difficult.
205: In contrast,
206: it turns out that calculation of the large deviation asymptotics of the optimally controlled network is quite feasible.
207: This is largely due to the fact that a {\it fixed} service policy invariably includes some {\it state discontinuities}.
208: For example,
209: a priority policy switches drastically when the highest priority queue empties.
210: When the policy is left as a parameter that is to be optimized these sharp discontinuities are not dealt with directly,
211: since the control and the large deviation behavior are identified simultaneously.
212: The situation is analogous to one found in the control of unconstrained processes such as diffusions.
213: If a fixed nonsmooth feedback control is considered then large deviation asymptotics are generally intractable,
214: but when the combined large deviation and optimal control problem is considered, much is possible \cite{dupmce}.
215:
216: For simplicity, we restrict in this paper to a class of Markovian networks,
217: and consider just one simple cost structure.
218: Much more general statistical models can be treated with similar arguments,
219: as can a more general cost.
220: A more fundamental restriction is on the routing in the network.
221: We assume a re-entrant line structure, so
222: that the input streams follow a fixed route through the network--we do not allow either randomized or controlled routing.
223: Relaxing the last conditions
224: leads to a problem that is significantly more difficult to analyze,
225: and would require a considerable extension of the results we prove.
226:
227: The deterministic game that is associated with the limit stochastic control
228: problem involves two players. One player
229: allocates service in a way analogous to the control in
230: the stochastic control problem, and the other player perturbs
231: the service and arrival rates. The cost is expressed in terms
232: of the large deviation rate function for the underlying arrival and service
233: processes, cumulated up to the time the dynamics exit the domain.
234: Heuristically, the first player
235: identifies those classes it is most worthwhile to allocate
236: service to, so as to delay the escape as much as possible and thereby maximize the cost.
237: The player who selects the perturbed rates attempts to
238: minimize the cost by driving it out of the domain,
239: while paying a cost for perturbing the rates.
240:
241: Our main result states that as the scaling parameter of the system
242: converges, the value for the stochastic control problem
243: converges to the value of the game. By way of proving the
244: result, we also show that the Hamilton-Jacobi-Bellman
245: equation associated with the game has a unique (Lipschitz continuous)
246: viscosity solution.
247:
248: Several works have considered
249: problems of optimal exit probabilities in the context of controlled
250: diffusion processes, in the asymptotically small white noise
251: intensity regime.
252: Fleming and Souganidis \cite{flesou} use viscosity
253: solutions techniques to study a controlled diffusion where the
254: control enters in the drift coefficient. Dupuis and Kushner \cite{dupkus}
255: extend their results to
256: the case where the diffusion coefficient is possibly degenerate.
257: Their technique relates the stochastic control problem
258: to the game in a more direct way, using time discretization,
259: without involving PDE analysis.
260: The stochastic control problem studied in the current paper
261: has the property that the jump rates in certain directions
262: (those that correspond to services, not to arrivals)
263: can be controlled to assume arbitrarily low values, including zero.
264: It appears to be a more subtle problem than the ones in the above
265: cited papers, in that it is analogous to a
266: controlled diffusion problem where the control
267: enters also in the {\em diffusion} coefficient, and where no uniform
268: non-degeneracy condition is assumed. This kind of degeneracy makes
269: it difficult to apply the time discretization idea of \cite{dupkus}.
270: The main idea of \cite{dupkus}, in which one directly relates
271: the control problem to the game, is still fruitful in the
272: current setting. Following this approach,
273: we relate the limit inferior [resp., superior] of the
274: asymptotic value for the control problem to the upper [resp., lower]
275: value of the game. However, showing that the game has value
276: and thereby obtaining the full convergence
277: result for the control problem requires a PDE analysis.
278:
279: The PDE analysis uses viscosity solutions methods.
280: There are three types of boundary conditions associated with the PDE:
281: Neumann, Dirichlet, and ``state space constraint.''
282: The first two types of boundary conditions correspond in the game
283: to the nonnegativity constraint on queue lengths and to stopping upon exit from the domain, respectively.
284: The third type of boundary condition arises when there are portions
285: of the boundary where exit can be blocked unilaterally by
286: one of the players, and it is optimal for it to do so.
287: It is well known since Soner \cite{son} that such a scenario
288: leads to the last boundary condition mentioned above.
289: Combining techniques of \cite{atadup2} and \cite{son},
290: we prove uniqueness of viscosity solutions for the PDE and show that
291: the game's upper and lower values are viscosity solutions,
292: thus establishing existence of value.
293: The trivial but crucial fact used in the uniqueness proof is
294: that the Isaacs condition holds (equation (\ref{eq:isaacs})).
295:
296: As an example, we analyze a case
297: where the domain is a hyper-rectangle,
298: and where the network consists of one server and many
299: queues, each customer requiring service only once.
300: We find an explicit solution to the corresponding PDE,
301: assuming the parameter $c$ is large enough.
302: This is only an initial result in this direction,
303: but it shows that explicit solutions can be found.
304: The solution turns out to be of particularly simple form
305: (see equation (\ref{eq:form})). The optimal service discipline
306: stemming from the solution corresponds to giving priority
307: to class $i$ whenever the state of the system is within
308: a subset $G_i$ of the domain. The partitioning of the domain
309: into subsets has a simple structure too (see Figure \ref{fig:example}
310: in Section \ref{sec:example} for an example in two dimensions).
311: See \cite{atadupshw} for explicit solutions in the case of tandem queues,
312: as well as identities relating the perturbed rates with
313: the unperturbed ones in a more general network.
314:
315: There is relatively little work on risk-sensitive and robust control of networks.
316: Ball et.\ al. have considered a robust formulation for network problems arising in vehicular traffic \cite{balday1,balday2},
317: and have explicitly identified the value function in certain instances.
318: Although their model is similar to ours in that the network dynamics are modeled via a Skorokhod Problem,
319: many other features, most notably the cost structure, are qualitatively different.
320: In addition,
321: the model they consider is not naturally related to a risk-sensitive control problem for a jump Markov model of a network.
322:
323:
324:
325: The organization of the paper is as follows.
326: Section \ref{sec:setting} introduces the network and the stochastic
327: control problem, describes a key tool in our analysis, namely
328: the Skorokhod Problem (SP), introduces the
329: differential game, and states the main result.
330: Section~\ref{sec:limit} establishes
331: the relation between the control problem asymptotics
332: and the game's upper and lower values.
333: Characterization of the upper and lower values of the game
334: as viscosity solutions of a PDE, as well as uniqueness for this PDE
335: are established in Section \ref{sec:pde}.
336: Section~\ref{sec:example} presents an example, and the paper concludes with Section~\ref{sec:lemmas}, which gives the proofs of several lemmas.
337: Throughout the paper,
338: numbering such as Lemma $a.b$ refers to the $b$th item of Lemma $a$.
339:
340:
341:
342:
343:
344: \section{Problem setting and the main result}\label{sec:setting}
345:
346: \noi\uu{\bf The queueing network control problem.}
347: We consider a system with $J$ customer classes, and without loss
348: assume that each class is identified with a queue at one of $K$ servers.
349: Each server provides service to at least one class.
350: Thus if $C(k)$ denotes the set of classes that are served by $k$,
351: then the control determines who receives service effort at server $k$
352: from among $i \in C(k)$. In particular, the sets $C(k)$, $k=1,\ldots, K$ are disjoint,
353: with $\cup_k C(k)=\{1,\ldots,J\}$.
354: The state of the network is the vector of queue lengths, denoted
355: by $X$.
356: After a customer of class $i$ is served, it turns into a customer of
357: class $r(i)$, where $i=0$ is used to denote the ``outside.''
358: We let $ e_j $ denote the unit vector in direction $j$ and set $ e_0 = 0 $
359: so that following service to class $j$ the state changes by
360: $e_{r(j)} - e_j$.
361: The control will be described by the vector $u=(u_1,...,u_J)$,
362: where $u_i=1$ if class $i$ customers are given service and $u_i=0$
363: otherwise. Since service can be given at any moment
364: to only one class at each station, the control vector must satisfy $\sum_{i\in C(k)}u_i\le1$ for
365: each $k$.
366: We next consider the scaled process $X^n$ under the scaling
367: which accelerates time by a factor of $n$ and shrinks space by
368: the same factor.
369: We are interested in a
370: risk-sensitive cost functional that is associated with
371: exit from a bounded set.
372: Let $G$ be a bounded subset of $\RJP $ that contains the origin
373: (additional assumptions on $G$ are given in Condition \ref{cond:G}).
374: Define
375: \[
376: \sigma^n \doteq \inf \{t:X^n(t) \not \in G \}.
377: \]
378: Then the control problem is to minimize the cost
379: $E_xe^{-nc\sigma^n}$,
380: where $E_x$ denotes expectation starting
381: from $x$, and $c>0$ is a constant.
382: With this cost structure ``risk-sensitivity'' means that atypically short exit times are weighted heavily by the cost.
383: A ``good'' control will avoid such an event with high
384: probability.
385: The significance from the point of view of stabilization of the
386: system is clear.
387: (See also \cite{dupmce} for the robust interpretation).
388:
389: A precise description of the stochastic control problem is as follows.
390: Let $G^n \doteq n^{-1}\Z_+^J \cap G$.
391: Define
392: $$
393: U \doteq \left\{(u_i), i=1,\ldots,J:\sum_{i\in C(k)} u_i\le1, k=1,\ldots,K,
394: u_i\ge0, i=1,\ldots,J\right\}.
395: $$
396: For $u\in U$ and $f:\Z^J\to\R$ let
397: \be\label{eq:gen_constr}
398: \til{\cal L}^{u}f(x)
399: \doteq \sum_{j=1}^{J}\lambda _{j}\left[ f(x+e_{j})-f(x)\right]
400: +\sum_{j=1}^J u_j \mu_j 1_{\{x+\tilde v_j\in\Z_+^J\}}
401: \lt[ f(x+\tilde v_j)-f(x)\right],
402: \ee
403: where $\til v_j=e_{r(j)}-e_j$.
404: It is assumed that for each $i$, $\la_i\ge0$,
405: while $\mu_i>0$.
406: For each $n\in\N$ consider the scaling defined by
407: \be\label{eq:gens}
408: \til\calL^{n,u}f(x) \doteq n\til\calL^ug(nx),
409: \ee
410: where $f:n^{-1}\Z^J\to\R$ and $g(\cdot)=f(n^{-1}\cdot)$.
411: A {\em controlled Markov process} starting from $x$ will consist of a
412: complete filtered
413: probability space $(\Om,\calF,(\calF_t),P_x^{n,u})$,
414: a state process $X^n$ taking values in $G^n$ that is continuous from the right
415: and with limits from the left,
416: a control process $u$ taking values in $U$,
417: such that $X^n$ is adapted to $\calF_t$,
418: $u$ is measurable and adapted to $\calF_t$,
419: $P_x^{u,n}(X^n(0)=x)=1$,
420: and for every function $f:n^{-1}\Z^J\to\R$
421: \[
422: f(X^n(t))-\int_0^t\til\calL^{n,u(s)}f(X^n(s))ds
423: \]
424: is an $\calF_t-$martingale.
425: $E_x^{n,u}$ denotes expectation with respect to $P_x^{n,u}$.
426: For a parameter $c>0$, the value function for the stochastic control problem is defined by
427: \be\label{eq:control}
428: V^n(x) \doteq -\inf n^{-1}\log E_x^{u,n}e^{-nc\sig_n},
429: \quad x\in G^n.
430: \ee
431: In this definition the infimum is over all controlled Markov processes.
432:
433: A measurable function $u(x,t)$, $u:G^n\times[0,\infty)\to U$
434: is said to be a {\em feedback control}.
435: We will make use of two well known facts:
436: to each feedback control there corresponds a controlled Markov process with $u(t)=u(X^n(t),t)$,
437: and in the definition of the value function the infimum can be restricted to feedback controls.
438:
439:
440: In the formulation just given we allow the maximizing player to choose a control from the convex set $U$.
441: This is a relaxed formulation,
442: which allows the server to simultaneously split the effort between 2 or more customer classes.
443: An alternative control space that is more natural in implementation consists of only the vertices of $U$,
444: in which case the server can only server one class at a time.
445: In a general game setting,
446: the distinction between such ``relaxed'' and ``pure'' control spaces can be significant.
447: However,
448: in the present setting it will turn out that the value is the same for both cases.
449: This is essentially due to the fact that the game arises from a risk-sensitive control problem,
450: which imposes additional structure on the game,
451: and will be further commented on below.
452:
453: \noi\uu{\bf Dynamics via the Skorokhod Problem.}
454: Our main goal will be to study the asymptotics of $V^n$,
455: and in particular, to show that they are
456: governed by the value of a deterministic differential
457: game. In order to define the dynamics of this game we first need a
458: formulation of the {\em Skorokhod Problem} (SP). We
459: give here the simplest formulation which covers our needs.
460: The reader is referred to \cite{dupram23} for a more general framework.
461: Let
462: \[
463: D_{+}([0,\infty ):\RJ ) \doteq \left \{\psi\in D([0,\infty ):\RJ ):
464: \psi(0)\in \RJP \right \},
465: \]
466: where $D([0,\infty ):\RJ )$ is the space of left continuous functions with right
467: hand limits, endowed with the uniform on compacts topology.
468: When restricting to continuous functions
469: we replace ``$D$'' with ``$C $''.
470: Let a set of vectors $\{ \gamma_i , \ i=1, \ldots , J \}$
471: be given and set $I(x) \doteq \{i:x_i=0\}$.
472: For each point $x$ on $ \partial \RJP $ -- the boundary of $\RJP $ -- let
473: \[
474: d(x) \doteq \left\{ \sum_{i \in I(x)}a_i \gamma_i: a_i \geq 0, \left\|
475: \sum_{i \in I(x)}a_i \gamma_i \right\| = 1 \right\}.
476: \]
477: The Skorokhod Map (SM) assigns to every path $\psi\in
478: D_{+ }([0,\infty ) :\RJ )$ a path $\phi$ that starts at $\phi(0)=\psi(0)$,
479: but is constrained to $\RJP $ as follows.
480: If $\phi$ is in the interior of $ \RJP $ then the evolution of $\phi$
481: mimics that of $\psi$,
482: in that the increments of the two functions are the same until $\phi$ hits
483: $ \partial \RJP $.
484: When $\phi$ is on the boundary a constraining ``force'' is applied to keep
485: $\phi$ in the domain,
486: and this force can only be applied in one of the directions $d(\phi(t))$,
487: and only for $t$ such that $\phi(t)$ is on the boundary.
488: The precise definition is as follows.
489: For $\eta\in
490: D([0,\infty ) : \RJ )$ and $t\in [0, \infty)$ we let $|\eta|(t)$ denote the total
491: variation of
492: $\eta$ on $[0,t]$ with respect to the Euclidean norm on $ \RJ $.
493:
494: \begin{definition}\label{def:sp}
495: Let $\psi\in D_{+}([0,\infty ) : \RJ )$ be given. Then $(\phi,\eta)$ solves
496: the SP for $\psi$ (with respect to $ \RJP $ and $\gamma_i, i=1,...,J $) if
497: $\phi(0)=\psi(0)$, and
498: if for all $t\in[0,\infty ) $
499: \begin{enumerate}
500: \item $\phi(t)=\psi(t)+\eta(t)$,
501: \item $\phi(t)\in \RJP $,
502: \item $|\eta|(t)<\infty$,
503: \item $|\eta|(t)=\int_{[0,t]} 1_{\{\phi(s)\in\partial \RJP \}}
504: d|\eta|(s)$,
505: \item There exists a Borel measurable function
506: $\gamma : [0,\infty ) \to \RJP $ such that $d|\eta|$-almost everywhere
507: $\gamma(t)\in d(\phi(t))$, and such that
508: $$ \eta(t) = \int_{[0,t]}\gamma(s)d|\eta|(s).
509: $$
510: \end{enumerate}
511: \end{definition}
512: Under a certain condition on $\{ \gamma_i \}$ (known in the literature as the
513: {\it completely}-${\cal S}$ condition \cite{reiwil}), it is known
514: that solutions to the SP exist in all of $D_{ +}([0,\infty): \RJP )$.
515: Under further conditions
516: (namely, {\em existence of the set $B$} -- see~\cite{harrei,dupish1,dupram23} and also Lemma \ref{lem:mu} below),
517: it is known that the Skorokhod Map is Lipschitz continuous, and consequently
518: the solution is unique.
519: Denoting the map $\psi\mapsto\phi$ in Definition \ref{def:sp} by
520: $\Gamma$, the Lipschitz property states that there is a constant
521: $K_1 $ such that
522: \be\label{sp:lip}
523: \sup_{t\in[0,\infty)}\|\Gamma(\psi_1)(t)-\Gamma(\psi_2)(t)\|
524: \le K_1 \sup_{t\in[0,\infty)}\|\psi_1(t)-\psi_2(t)\|,\quad
525: \psi_1,\psi_2\in D_{ + }([0,\infty): \RJ ).
526: \ee
527: The SP that will be considered here is the one for which
528: $\gamma_i=e_i-e_{r(i)} = - {\tilde v}_i $. For this problem, the following is well known.
529: \begin{theorem}[\cite{harrei,dupram23}]\label{th:SP}
530: The SP associated with the domain $ \RJP $ and the constraint vectors
531: $\gamma_i, i=1,...,J $
532: possesses a unique solution, and the Skorokhod Map is
533: Lipschitz continuous on the space $D_{ +}([0,\infty): \RJP )$.
534: Moreover, the Skorokhod Map takes
535: $ C_{+}([0,\infty): \RJ ) $ into $ C_{+}([0,\infty): \RJ ) $,
536: and therefore $ \Gamma (\phi ) $ is continuous if $ \phi $ is.
537: \end{theorem}
538: We next define a constrained ordinary differential equation.
539: As is proved (in greater generality) in \cite{dupish1},
540: one can define a projection $\pi: \RJ \goto \RJP $ that is consistent with
541: the constraint directions $\{\gamma_i, i=1,...,J \}$,
542: in that $\pi(x) = x$ if $x \in \RJP $,
543: and if $x \not \in \RJP $ then $\pi(x)-x = \alpha r$,
544: where $\alpha \geq 0$ and $r \in d(\pi(x))$.
545: With this projection given,
546: we can now define
547: for each point $x \in \partial \RJP $ and each $v \in \RJ $
548: the {\it projected velocity}
549: \[
550: \pi(x,v) \doteq \lim_{\Delta \downarrow 0} \frac{\pi(x+\Delta
551: v)-\pi(x)}{\Delta}.
552: \]
553: For details on why this limit is always well defined and further properties
554: of the projected velocity we refer to \cite[Section 3 and Lemma
555: 3.8]{buddup}.
556: Let $v:[0,\infty) \goto \RJ $ have the property that each component of
557: $v$ is integrable over each interval $[0,T]$, $T<\infty$.
558: Then the ODE of interest takes the form
559: \be
560: \label{eqn:ODE}
561: \dot \phi(t) = \pi(\phi(t),v(t)), \quad \phi (0) = \phi_0 \in \RJP .
562: \ee
563: An absolutely continuous function $\phi:[0,\infty) \goto \RJP $ is a solution
564: to (\ref{eqn:ODE}) if the equation is satisfied a.e.\ in $t$.
565: By using the regularity properties (\ref{sp:lip}) of the associated Skorokhod Map
566: and because of the
567: particularly simple nature of the right hand side,
568: one can show that $\phi$ solves (\ref{eqn:ODE}) if and only if $\phi$ is
569: the image of $\psi(t) \doteq \int_0^t v(s)ds +x$ under the Skorokhod
570: Map, and thus all the standard qualitative properties (existence and
571: uniqueness of solutions, stability with respect to perturbations, etc.)
572: hold \cite{dupish1,dupnag}.
573:
574: As mentioned above, the SP formulation will be our means of defining
575: the dynamics of a deterministic game. Before discussing this game,
576: let us show how the same formulation is also useful for
577: the stochastic control problem defined earlier in this section.
578: First, since $\til v_j = e_{r(j)}-e_j = - \gamma_j$, it is easy to verify that for the particular SP considered here
579: $\pi(x,v)=v1_{x+v\in\Z_+^J}$ for all $x\in\Z_+^J$ and
580: $v\in\{\til v_j:j=1,\ldots,J\}$.
581: Therefore the generator $\til\calL^{u}$ of (\ref{eq:gen_constr})
582: can also be written as
583: $$
584: \til\calL^uf(x)=\sum_{j=1}^J\la_j[f(x+e_j)-f(x)]+
585: \sum_{j=1}^Ju_j\mu_j[f(x+\pi(x,\tilde v_j))-f(x)].
586: $$
587: A measurable function $u(t)$, $u:[0,\infty)\to U$
588: is said to be an {\em open loop control}.
589: Note that this control has no state feedback.
590: When $u$ is an open loop control,
591: it is possible to write the corresponding controlled process $X$ as $\Gamma(Y)$.
592: The process $Y$, which will be called
593: {\em the unconstrained controlled process},
594: is a controlled Markov process with a simpler structure.
595: To be precise, let
596: $$
597: \til{\cal L}_0^{u}f(x)
598: =\sum_{j=1}^{J}\lambda _{j}\left[ f(x+e_{j})-f(x)\right]
599: +\sum_{j=1}^Ju_j\mu_j\left[ f(x+ \tilde v_j)-f(x)\right],
600: $$
601: and let $\til\calL_0^{n,u}$ be defined analogously to (\ref{eq:gens}).
602: A controlled Markov process
603: $X^n$ on $G^n$ [respectively, $Y^n$ on $n^{-1}\Z^J$] is defined as before,
604: but now using the
605: generator $(\calL^{n,u}f)(t)=(\til\calL^{n,u(t)}f)(x)$
606: [resp., $(\calL_0^{n,u}f)(t)=(\til\calL_0^{n,u(t)}f)(x)$].
607: The simplification that the SP introduces
608: is that if $u$ is an open loop control, and if $Y^n$ is a
609: controlled Markov process corresponding to $\calL_0^{n,u}$
610: on $(\Om,\calF,(\calF_t),P_x^{n,u})$, then $X^n=\Gamma(Y^n)$
611: is a controlled Markov process corresponding to $\calL^{n,u}$
612: on the same filtered probability space.
613: The role played by the SP in relating constrained and
614: unconstrained processes is exhibited here in a simple fashion,
615: for introductory purposes. We will, in fact, use it in
616: a slightly more complicated setting later on in Lemma \ref{lem:xbar} and Lemma
617: \ref{lem:xbar2}.
618:
619: \noi\uu{\bf A differential game.}
620: In this paper,
621: we prove that the value function $V^n(x)$ for a stochastic control problem associated with our queueing network model is approximately equal (for large $n$) to the value function of a related differential
622: game.
623: In addition,
624: the dynamics of this game are defined in terms of an associated SP.
625: Before introducing the game
626: formally we explain why this is to be expected.
627: In a problem with no control, the exponential decay rate
628: of quantities such as $Ee^{-cn\sig_n}$ is
629: given in terms of the sample path
630: large deviation rate function associated with the process, which in
631: turn can be expressed in terms of the rate function for the Poisson
632: primitives that drive the model.
633: This is supported by the
634: well known Laplace's principle \cite{dupell}. Heuristically,
635: one thinks of the rate function as a cost paid for changing the measure
636: so as to make the rare event of exiting
637: on short time interval a probable event.
638: Laplace's principle asserts that the decay rate can be expressed
639: as the solution to a deterministic optimization problem
640: involving the cost $-c\sig$ combined with the cost of changing the measure:
641: cf.\ \cite[Eq.\ (5.20)--(5.23)]{shwe}.
642: When the stochastic model involves optimal control, there is one
643: more variable to optimize over in the limit, and this results in a game.
644: The game's deterministic dynamics are the natural law of large numbers limit under the changed measure.
645: Boundary constraints and constraining meachanisms which are present in the prelimit model are represented in the limit model by the SP.
646: The cost for the game involves
647: the large deviation rate function for the Poisson primitives,
648: the time till the dynamics exits the domain, and the parameter $c$.
649:
650: We thus consider a zero sum game involving two players.
651: One (which we call the maximizing player)
652: selects the service allocation and attempts to maximize.
653: The other (called the minimizing player)
654: chooses the perturbed arrival and service rates and attempts to minimize.
655: Throughout, the perturbed rates will be denoted by an overbar,
656: as in $\bar\la_i,\bar\mu_i$.
657:
658: Recall that for $u\in U$, $u_i$
659: stands for the fraction of service effort given to class $i$.
660: The control space for the maximizing player is
661: $$ \bar U\doteq \{u:[0,\infty)\to U\ ;\ u \mbox{ is measurable}\}.
662: $$
663: Let $l:\R\to\R_+\cup\{+\infty\}$
664: be defined as
665: $$
666: l(x)\doteq \lt\{\begin{array}{ll}x\log x-x+1 & x\ge0, \\ +\infty & x<0,
667: \end{array}\rt.
668: $$
669: where $0\log0\doteq0$.
670: Denoting $M =[0,\infty)^{2J}$, the control space for the minimizing
671: player will be
672: \begin{equation}
673: \label{embar}
674: \bar M =\{m=(\bar\la_1,\ldots,\bar\la_J,\bar\mu_1,\ldots,\bar\mu_J):
675: [0,\infty)\to M;
676: \ \mbox{$m$ is measurable, $l\circ m$ is locally integrable}\}.
677: \end{equation}
678: For $u\in U$ and $m\in M$ define
679: \[
680: v(u,m) \doteq \sum_{j=1}^J \bar \lambda_jv_j + \sum_{i=1}^J
681: u_i \bar\mu_i \tilde v_i,
682: \]
683: where $v_j=e_j$, and as before $\tilde v_i=e_{r(i)}-e_i$.
684: Then the dynamics are given by
685: \[
686: \lt\{\begin{array}{ll} \dot \phi(t) = \pi (\phi(t),v(u(t),m(t))), \\
687: \phi(0)=x.
688: \end{array}\rt.
689: \]
690: To define the cost for the game, let
691: $\rho:U\times M\to\R_+\cup\{+\infty\}$ be
692: $$ \rho(u,m)\doteq \sum_{i=1}^{J}\lambda _{i}l\left( \frac{\bar{\lambda}_{i}}{
693: \lambda _{i}}\right) +\sum_{i=1}^{J}u_i\mu _il\left( \frac{\bar{\mu}_i}
694: {\mu _i}\right).
695: $$
696: By convention, if $\la_i=0$ and $\bar\la_i>0$ for some $i$, we let
697: $\rho=\infty$ (recall that by assumption, $\mu_i>0$).
698: Let the exit time be defined by
699: $$
700: \sig\doteq \inf\{t:\phi(t)\not\in G\}.
701: $$
702: With $c > 0$ as in~(\ref{eq:control}),
703: the cost is given by
704: $$ C(x,u,m)=\int_0^\sig [c+\rho(u(t),m(t))]dt.
705: $$
706:
707: As in \cite{ellkal} we need the notion
708: of strategies. We endow both $\bar U$ and $\bar M$ with the metric
709: $\til\rho(\om_1,\om_2)=\sum_n2^{-n}(\int_0^n
710: |\om_1(t)-\om_2(t)|dt \w1)$,
711: and with the corresponding Borel $\s$-fields.
712: A mapping $\al:\bar M \to\bar U$ is called a
713: {\em strategy for the maximizing player}
714: if it is measurable
715: and if for every $m,\tilde m\in\bar M$ and $t>0$ such that
716: $$ m(s)=\tilde m(s) \mbox{ for a.e.\ } s\in[0,t],
717: $$
718: one has
719: $$ \al[m](s) = \al[\tilde m](s) \mbox{ for a.e.\ } s\in[0,t].
720: $$
721: In an analogous way, one defines a mapping $\beta:\bar U\to \bar M$
722: to be a {\em strategy for the minimizing player}.
723: The set of all strategies for the maximizing [resp., minimizing]
724: player will be denoted by $A$ [resp., $B$].
725: The lower value for the game is defined as
726: \[
727: V^-(x) = \inf_{\beta\in B} \sup_{u\in\bar U}C(x,u,\beta[u]),
728: \]
729: and the upper value as
730: $$
731: V^+(x)=\sup_{\al\in A}\inf_{m\in\bar M}C(x,\al[m],m).
732: $$
733: To avoid confusion, we remark that despite the terms ``upper'' and
734: ``lower'' value, it is not in general obvious that $V^-\le V^+$.
735:
736:
737: \noi\uu{\bf Main result.}
738: We make the following assumption on the domain $G$. Let
739: $$
740: \calJ_+\doteq\{i\in\{1,\ldots,J\}:\la_i>0\}.
741: $$
742:
743: \begin{condition}\label{cond:G}
744: We assume that the domain $G$ satisfies one of the following.
745: \begin{enumerate}
746: \item
747: $G$ is a rectangle given by
748: $$ G=\{(x_1,\ldots,x_J): 0\le x_i<z_i, i\in\calJ_+;\ 0\le x_j\le z_j,
749: j\not\in\calJ_+\},
750: $$
751: for some $z_i>0$, $i=1,\ldots,J$.
752: \item
753: $G$ is simply connected and bounded, and given by
754: $$
755: G=\bigcap_{i\in\calJ_+}G_i,
756: $$
757: where for $i\in\calJ_+$, we are given positive Lipschitz functions
758: $\phi_i:\R^{J-1}\to\R$, and
759: $$
760: G_i=\{(x_1,\ldots,x_J)\in \RJP : 0\le x_i<\phi_i(x_1,\ldots,x_{i-1},
761: x_{i+1},\ldots,x_J)\}.
762: $$
763: \end{enumerate}
764: \end{condition}
765: This condition covers many typical constraints one would consider on buffer size,
766: including separate constraints on individual queues (Condition \ref{cond:G}.1) and one constraint on the sum of the queues (Condition \ref{cond:G}.2).
767:
768: The shape of the domain is simpler in Condition \ref{cond:G}.1,
769: in that it is restricted to a hyper-rectangle.
770: On the other hand, it is also possible under this condition for the
771: maximizing player to unilaterally prevent an exit
772: through a certain portion of $\pl G\setminus\pl\R_+^J$.
773: Although it is in principle possible that the dynamics
774: could exit through this portion of the boundary, it will always be optimal for
775: the maximizing player to not allow it.
776: Consider the simple network illustrated in Figure \ref{fig1}.
777: The maximizing player can prevent exit through the dashed portion of the boundary simply by stopping service at the first queue.
778: As a consequence,
779: there are in general three different types of boundary--the constraining boundary due to non-negativity constraints on queue length,
780: the part of the boundary where exit can be blocked,
781: and the remainder.
782: These three types of boundary behavior will result,
783: in the PDE analysis, in three types of boundary conditions.
784: We now define the three portions of the boundary.
785: Under Condition \ref{cond:G}.1, let
786: $$
787: \pl_cG=\{(x_1,\ldots,x_J)\in G: x_j=z_j,\,\mbox{some}\,j\not\in\calJ_+\}.
788: $$
789: For notational convenience, we let $\pl_cG=\emptyset$ under
790: Condition \ref{cond:G}.2. In both cases we then set
791: $$
792: \pl_oG=\pl G\setminus G, \qquad
793: \pl_+G=(G\cap\pl\R_+^J)\setminus\pl_cG.
794: $$
795: Note that in both cases, $\pl_cG$, $\pl_oG$ and $\pl_+G$
796: partition $\pl G$. Also, $\pl_cG\subset G$ while $\pl_oG\subset
797: G^c$. As usual, we will denote $G^o=G\setminus\pl G$
798: and $\bar G=G\cup\pl G$.
799: $ \pl_cG $ is the part of the boundary were the
800: maximizing player can prevent the dynamics from exiting,
801: and $ \pl_oG $ is the part where it can not.
802: Finally, it will be convenient to denote
803: $$
804: \pl_{co}G=\pl_oG\cup\pl_cG.
805: $$
806:
807:
808: \begin{figure} % export: 30 percent.
809: \centerline{
810: \begin{tabular}{cc}
811: \psfig{file=fig_net_1.eps}\qquad\qquad\qquad
812: \psfig{file=fig_ex_1.eps}
813: \end{tabular}
814: }
815: \caption{
816: A simple queueing network, a rectangular domain and three
817: types of boundary.
818: Full line: $~$ \hfill $\pl_+G$, dashed line: $\pl_cG$,
819: and dotted line: $\pl_oG$
820: \hfill $~$ }
821: \label{fig1}
822: \end{figure}
823:
824:
825: Our main result is the following.
826: \begin{theorem}\label{th:main}
827: Let Condition \ref{cond:G} hold.
828: Then $V^+=V^-\doteq V$ on $G$. Moreover, if $x_n\in G^n$,
829: $n\in\N$ are such that $x_n\to x\in G$, then
830: $\lim_{n\to\infty}V^n(x_n) = V(x)$.
831: \end{theorem}
832:
833: \remark
834: A stronger form of the convergence statement in fact holds.
835: Namely,
836: $$
837: \limsup_{\eps\downarrow0}
838: \limsup_{n\to\infty}\sup\{|V^n(x)-V(y)|: x\in G^n,y\in G, |x-y|\le\eps\}
839: =0.
840: $$
841: This is an immediate consequence of Theorem \ref{th:main}
842: and the fact that for each $n$, $V^n$ is Lipschitz on $G^n$,
843: with a constant that does not depend on $n$ (Lemma \ref{lem:unif}).
844: \qed
845:
846: The proof is established in two major steps.
847: Step 1 will be an immediate consequence of the main results of
848: Section \ref{sec:limit},
849: and Step 2 will follow from Section \ref{sec:pde}.
850:
851: \noi{\em Step 1.} We define a version of the game, technically
852: easier to work with, in which
853: all perturbed rates ($\bar\la_i,\bar\mu_i$) are bounded
854: by $b<\infty$. The corresponding upper and lower values,
855: defined analogously, are denoted by $V^{b,+}$ and $V^{b,-}$.
856: Then we show that for all $b$ large enough
857: (cf.\ Theorem \ref{th:limit})
858: $$
859: V^{b,+} (x)\le\liminf_{n\to\infty}V^n(x_n)\le\limsup_{n\to\infty}V^n(x_n)
860: \le V^{b,-} (x) .
861: $$
862:
863: \noi{\em Step 2.}
864: We show that for $b$ large,
865: $V^{b,+} =V^{b,-} $ on $G$. To this end, we formulate a PDE
866: for which we show that uniqueness of (Lipschitz) viscosity solutions
867: holds (Theorem \ref{th:unique}),
868: and also show that both $V^{b,+}$ and $V^{b,-}$ are viscosity solutions
869: (Theorem \ref{th:solve}).
870: Since $ V^n (x) $ does not depend on $b$, neither do $ V^{b,\pm} (x) $.
871: Theorem \ref{th:main} follows.
872:
873:
874: \section{The control problem and the game}\label{sec:limit}
875:
876: We begin by stating some basic properties of the stochastic
877: control problem and of the deterministic game.
878: The proofs of these properties are
879: deferred to Section \ref{sec:lemmas}.
880:
881: Consider the following generators, defined for any $u\in U$ and
882: $m\in M$, for constrained and unconstrained controlled Markov
883: processes:
884: \beaa
885: \calL^{n,u,m}f(x) &=& \sum_{j=1}^J n \bar \lambda_j \left [f \left (x+\frac{1}{n} v_j \right )-f(x) \right ]
886: + \sum_{i=1}^J n \bar \mu_i u_i\left [ f \left ( x+\frac{1}{n}\pi(x, \tilde v_i) \right )-f(x) \right ], \\
887: \calL_0^{n,u,m}f(x)
888: &=& \sum_{j=1}^J n \bar \lambda_j \left [ f \left (x+\frac{1}{n} v_j \right )-f(x) \right ]
889: + \sum_{i=1}^J n \bar \mu_i u_i \left [f \left ( x+\frac{1}{n} \tilde v_i \right )-f(x) \right ].
890: \eeaa
891: The definition of the corresponding controlled processes will be made precise in Lemmas
892: \ref{lem:xbar} and \ref{lem:xbar2}.
893:
894: Owing to the logarithmic transform in (\ref{eq:control}), one expects $V^n$ to satisfy an Isaacs equation \cite{flesou}.
895: In fact, $V^n$ satisfies
896: \be\label{eq:dpe}
897: \lt\{\begin{array}{cc}
898: 0=\sup_{u\in U} \inf_{m\in M} [\calL^{n,u,m}V^n(x) +c+\rho(u,m)],
899: & x \in G^n \\
900: V^n(x)=0, & x \not \in G^n.
901: \end{array}\rt.
902: \ee
903: We comment that this is also the dynamic programming equation (DPE)
904: for an associated stochastic game
905: that is related to the deterministic game
906: via a law of large numbers scaling and limit,
907: and will not be further considered in this paper.
908:
909:
910: \begin{lemma}\label{lem:dpe}
911: The value function $V^n$ of (\ref{eq:control}) uniquely solves
912: the DPE (\ref{eq:dpe}).
913: \end{lemma}
914:
915: The following lemma gives a key estimate on the value function.
916:
917: \begin{lemma}\label{lem:unif}
918: Under Condition~\ref{cond:G},
919: $ V^n (x) $ satisfies the Lipschitz property
920: on $(n^{-1}\Z_+^J) \cap \bar G$
921: with a constant that does not depend on $n\in\N$.
922: Consequently, $ \sup_{n, x \in G^n} V^n (x) < \infty $.
923: \end{lemma}
924: We comment that the above result is, in general, not valid for $V^n$
925: on $n^{-1}\Z_+^J$, since $V^n$ changes abruptly near
926: the portion $\pl_cG$ of the boundary.
927:
928: For each fixed $u\in U$, the mapping $m\rightarrow \rho(u,m)$,
929: when restricted to $\bar \mu_i$ such that $u_i>0$, is strictly convex with compact level sets.
930: We conclude that the infimum over $m$ in the DPE
931: is achieved,
932: and denote such a point by $m^n(x,u)$.
933: Although part 1 of the following lemma is not used elsewhere,
934: it indicates why the Isaacs condition should hold in (\ref{eq:dpe}).
935:
936: \begin{lemma}\label{lem:bound:m} Let Condition \ref{cond:G} hold.
937: Then
938: \ben
939: \item
940: $m^n(x,u)$ can be chosen independently of $u$, and
941: \item
942: there is $b_0<\infty$ such that for all $x,n$ and $u$, $ m^n(x,u) \leq b_0$.
943: \een
944: \end{lemma}
945:
946:
947: We introduce two parametric variations of the game defined in
948: Section \ref{sec:setting}. The first will be associated with
949: domain perturbation (parameterized by the symbol $a$), and the second
950: with a bound on the perturbed rates (parameterized by the symbol $b$).
951:
952: For some fixed $a_0>0$, consider perturbations $G_a$,
953: $a\in(-a_0,a_0)$ of the domain $G$
954: defined as follows. If $G$ satisfies Condition~\ref{cond:G}.1,
955: then $G_a$ is defined as $G$, but with $z_i$ replaced by $z_i+a$,
956: $i=1,\ldots,J$.
957: If $G$ is as in Condition~\ref{cond:G}.2, then $G_a$ is defined
958: as $G$, but where $\phi_i$ is replaced by $\phi_i+a$, $i\in\calJ_+$.
959:
960: For any $b\in(0,\infty)$, let $M^b=[0,b]^{2J}$. Analogously to
961: the definition (\ref{embar}) of $\bar M$, let
962: \begin{equation}
963: \label{embarb}
964: \bar M^b =\{m=(\bar\la_1,\ldots,\bar\la_J,\bar\mu_1,\ldots,\bar\mu_J):
965: [0,\infty)\to M^b \ ;\ \mbox{$m$ is measurable}\}.
966: \end{equation}
967: Strategies and values for the game are then defined
968: analogously to the way strategies and values are defined for
969: the original game, using $\bar M^b$ in place of $\bar M$.
970: It will be convenient to set $M^\infty\doteq M$ and
971: $\bar M^\infty\doteq\bar M$, and to refer to the original game
972: of Section \ref{sec:setting} as the case $b=\infty$.
973:
974: The cost, sets of strategies, lower and upper values of the
975: games resulting by the introduction of the parameters
976: $a$ and $b$ will be denoted as $C_a(x,u,m)$, $A^b$, $B^b$,
977: $V^{b,-}_a$ and $V^{b,+}_a$. When $a=0$ [resp., $b=\infty$],
978: the dependence on $a$ [$b$] will be eliminated from the notation,
979: as in $V^-_a$, $V^{b,-}$.
980:
981: Let $b_0$ be as in Lemma \ref{lem:bound:m}.
982: Denote
983: \begin{equation}
984: \label{eq:lal}
985: b^* \doteq \max \{ b_0, \lambda_i , \mu_i , i = 1 , \ldots , J \} + 1.
986: \end{equation}
987:
988: \begin{lemma}\label{lem:cont}
989: Assume Condition~\ref{cond:G}.
990: Then
991: \ben
992: \item
993: $\dist(\pl_{co} G_a,\pl_{co} G)\doteq
994: \inf\{|x-y|:x \in \pl_{co} G_a, y \in \pl_{co} G\} >0\mbox{ if } 0<|a|<a_0$;
995: \item
996: the values
997: $V^{b,\pm}$ are bounded on $G$, uniformly for $b\in[b^*,\infty]$,
998: and there is a constant $c_0$ such that for any
999: $x\in G$, $|a|<\eps$ (where $\eps$ depends on $x$), and $b\in[b^*,\infty]$,
1000: one has
1001: $|V^{b,-}_a(x)-V^{b,-}(x)|\le c_0|a|$ and
1002: $|V^{b,+}_a(x)-V^{b,+}(x)|\le c_0|a|$.
1003: \een
1004: \end{lemma}
1005:
1006:
1007: The following lemma shows that any nearly optimal strategy for the minimizing player will satisfy a uniform upper bound on the integrated running cost.
1008: Moreover, there is a finite time $T_0$ such that
1009: for each such minimizing strategy, any open loop control used by the maximizing player leads to exit by $T_0$.
1010: Similarly,
1011: given any strategy for the maximizing player the minimizing player can restrict to open loop controls that force exit by $T_0$.
1012: \begin{lemma}\label{lem:apriori}
1013: Fix $b\in[b^*,\infty]$.
1014: Given $\beta \in B^b$,
1015: write
1016: $(\bar\la_i(\cdot ),\bar\mu_i(\cdot ))=\beta[u](\cdot)$. For $z,T>0$ let
1017: $B^{z,T}$ denote the set of $\beta\in B^b$ which satisfy
1018: $$
1019: \int_0^T\sum_i[\la_il(\bar\la_i(t)/\la_i)+u_i(t)
1020: \mu_il(\bar\mu_i(t)/\mu_i)]dt
1021: \le z,
1022: $$
1023: for all $u\in\bar U$.
1024: For $\al\in A^b$, let $\bar M(\al,T)$ denote the set of $m\in\bar M$
1025: for which $\s(x,\al[m],m)\le T$.
1026: Then there are constants $z_0,T_0>0$ such that
1027: $$
1028: V^- (x)=\inf_{\beta\in B^{z_0,T_0}}\sup_{u\in\bar U}
1029: \int_0^{\sig\w T_0}[c+\rho(u(t),\beta[u](t))]dt,
1030: $$
1031: and
1032: $$
1033: V^+(x)=\sup_{\al\in A}\inf_{m\in\bar M(\al,T_0)}
1034: \int_0^{\sig\w T_0}[c+\rho(\al[m](t),m(t))]dt.
1035: $$
1036: \end{lemma}
1037:
1038: In the rest of the section the strategies $\beta$
1039: will be assumed (without loss)
1040: to be in $B^{z_0,T_0}$, where $z_0,T_0$ are as in
1041: Lemma \ref{lem:apriori}, and are fixed throughout. Also, $m\in\bar M$
1042: will be assumed to be in $\bar M(\al,T_0)$ whenever it is
1043: clear which $\al$ is considered.
1044: With an abuse of notation, we denote
1045: $B^{z_0,T_0}$ by $B$.
1046:
1047: \begin{lemma}\label{lem:vallip} Under Condition \ref{cond:G},
1048: $V^{b,-}$ and $V^{b,+}$ are Lipschitz on $G$, uniformly for
1049: $ b\in[b^*,\infty]$.
1050: \end{lemma}
1051:
1052: We are now ready to prove the following result.
1053: \begin{theorem}\label{th:limit}
1054: Let Condition \ref{cond:G} hold, and let $b\in[b^*,\infty)$.
1055: Then for any $x\in G$ and any $\{ x_n \} $ converging to $x$ (with
1056: $x_n\in G^n$),
1057: $$
1058: V^{b,+}(x)\le\liminf_{n\to\infty}V^n(x_n)\le\limsup_{n\to\infty}
1059: V^n(x_n)\le V^{b,-}(x).
1060: $$
1061: \end{theorem}
1062:
1063: \proofOf{Theorem~\ref{th:limit}}
1064: The result is established by considering a sequence of stochastic
1065: processes, defined using the constrained ODEs, but for which
1066: the controls $u$ and $m$ are governed by, on one hand, a nearly optimal
1067: strategy for the game, and on the other hand, a nearly optimal
1068: control for the stochastic control problem.
1069: The technique uses standard martingale estimates, and is based
1070: on the construction (deferred to Section \ref{sec:lemmas})
1071: of an auxiliary controlled Markov process
1072: that is controlled by
1073: the selected strategy and stochastic control.
1074:
1075: \noi{\em \uu{Upper bound}}
1076:
1077: \noi
1078: Fix $b\in[b^*,\infty)$.
1079: The dependence on $b$ will be suppressed in the notation for $V^-$, $V^-_a$, etc.
1080: We first show that
1081: \be\label{eq:ub}
1082: \limsup_{n\to\infty} V^n(x_n) \leq V^-(x).
1083: \ee
1084: According to Lemma~\ref{lem:cont}.2, it is enough to show that
1085: for all $a>0$
1086: \[
1087: \limsup_{n\to\infty} V^n(x_n) \leq V_a^-(x).
1088: \]
1089: Let $\beta\in B^b $ and $a>0$ be given,
1090: and set $C_a(x,\beta)=\sup_{u\in\bar U} C_a(x,u,\beta[u])$.
1091: It is enough to show that
1092: \be\label{eq:ca}
1093: \limsup_{n\to\infty} V^n(x_n) \leq C_a(x,\beta), \quad a>0.
1094: \ee
1095: We therefore fix $\beta$ throughout, and turn to prove (\ref{eq:ca}).
1096: We can assume without loss that
1097: \begin{equation}\label{eq:31}
1098: C_a(x,\beta)<\infty.
1099: \end{equation}
1100: Note that in the DPE (\ref{eq:dpe}) the supremum is with respect to $u$ in
1101: a compact set $U$, and that the function being maximized is continuous in $u$ (for each $y$).
1102: Let $u^n(y)$ denote a point where it is achieved.
1103: Then for any $m\in M$ and $y\in G^n$,
1104: \be\label{ineq1}
1105: 0 \leq \calL^{n,u^n(y),m}V^n(y) +c+\rho(u^n(y),m).
1106: \ee
1107: \begin{lemma}\label{lem:xbar}
1108: Let $n$ be fixed, and let $b\in[b^*,\infty)$,
1109: $\beta$ and $x_n$ be as above.
1110: Then there is a filtered probability space
1111: $(\bar\Om,\bar F,(\bar F_t),\bar P)$, and $\bar F_t$-adapted
1112: RCLL processes
1113: $\bar X^n$, $\bar Y^n$ and $m^n$ such that with $\bar P$-probability one
1114: $m^n(t)=\beta[\bar u^n](t)$ a.e.\ $t$, $\bar u^n(t)=u^n(\bar X^n(t))$,
1115: $\bar X^n=\Gamma(\bar Y^n)$,
1116: $\bar X^n(0)=\bar Y^n(0)=x_n$,
1117: and for any $f$
1118: \[
1119: f(\bar X^n(t)) - \int_0^t \calL^{n,\bar u^n, m^n(s)}f(\bar X^n(s))ds
1120: \]
1121: \[
1122: f(\bar Y^n(t)) - \int_0^t \calL_0^{n,\bar u^n, m^n(s)}f(\bar Y^n(s))ds
1123: \]
1124: are $(F_t)$-martingales.
1125: Moreover, with $T_0$ as in Lemma~\ref{lem:apriori} let $N_n $ denote
1126: the total number of jumps of $ \bar Y^n $ on $ [0,T_0]$. Then
1127: \be\label{ineq:Nn}
1128: EN_n\le 2JT_0bn.
1129: \ee
1130: \end{lemma}
1131:
1132: \proof See Section \ref{sec:lemmas}.
1133:
1134: Returning to the proof of~(\ref{eq:ca}), let $\bar u^n(t) \doteq u^n(\bar X^n(t))$
1135: and let $\bar\sigma^n$ be the first
1136: exit time of $\bar X^n$ from $G$.
1137: Combining
1138: (\ref{ineq1}) and Lemma \ref{lem:xbar}, for any bounded stopping
1139: time $S\le\bar\sig^n$,
1140: \be\label{ineq2}
1141: V^n(x_n) \leq \bar E_{x_n} \left[ V^n(\bar X^n(S))
1142: +\int_0^{S}[c+\rho(\bar u^n(s),\beta[\bar u^n](s))]ds \right].
1143: \ee
1144: Denoting $\beta[\bar u^n](t)= \{ (\bar\la_i^n(t),\bar\mu_i^n(t)) \} $,
1145: define $\phi^n$ as $\phi^n=\Gamma(\psi^n)$, where
1146: \[
1147: \psi^n (t)= x+\int_0^t v(\bar u^n,\beta[\bar u^n])ds,
1148: \]
1149: and let
1150: \[
1151: \hat\sig^n_a \doteq \inf\{t:\phi^n (t) \not \in G_a\}.
1152: \]
1153: Then the definition of $C_a(x,\beta)$ implies
1154: \[
1155: \int_0^{\hat\sig^n_a}[c+\rho(\bar u^n(s),\beta[\bar u^n](s))]ds
1156: \leq C_a(x,\beta).
1157: \]
1158: Apply (\ref{ineq2}) with $S=\hat\sig_a^n\w\bar\sig^n\w T$.
1159: If $T$ is sufficiently large, then (\ref{eq:31}) and the
1160: fact that $c>0$ imply $\hat\sig_a^n \leq T$. Thus, using
1161: $\bar E_{x_n}V^n(\bar X^n(\bar\sig^n)))=0$,
1162: \[
1163: V^n(x_n) \leq \bar E_{x_n} \left[ V^n(\bar X^n(\hat\sig^n_a))
1164: 1_{\{\hat\sig^n_a < \bar\sigma^n\}}
1165: +\int_0^{\hat\sig^n_a}[c+\rho(\bar u^n(s),\beta[\bar u^n](s))]ds \right].
1166: \]
1167: Again using the uniform boundedness of $V^n(x)$ (Lemma \ref{lem:unif}), there is a constant $b_2 <\infty$
1168: such that for all $n$
1169: \be\label{ineq3}
1170: V^n(x_n) \leq b_2 \bar P_{x_n}(\hat\sig^n_a \le \bar\sigma^n)
1171: + C_a(x,\beta).
1172: \ee
1173: In what follows, we show that $\bar P_{x_n}(\hat\sig^n_a \le \bar\sigma^n)$
1174: tends to zero. To this end,
1175: note that $\calL_0^{n,u,m}\,{\rm id}(y)=\sum_i\bar\la_iv_i+\sum_iu_i\bar\mu_i
1176: \tilde v_i$, where ${\rm id}$ is the identity map.
1177: Therefore, using again Lemma \ref{lem:xbar},
1178: \[
1179: \bar Y^n(t) -x_n = \int_0^t \lt[ \sum_{i=1}^J \bar\la_i^n(s)v_i
1180: +\sum_{i=1}^J\bar u_i^n(s)\bar\mu_i^n(s)\tilde v_i\rt]
1181: ds + \eta^n(t),
1182: \]
1183: where $\eta^n$ is a zero mean martingale.
1184: To prove that
1185: \be\label{conv:mgale}
1186: \sup_{t\in[0,T]}|\eta^n(t)| \to 0 \quad \mbox{in distribution},
1187: \ee
1188: it is enough, by Doob's maximal inequality, to show that
1189: $$
1190: \bar E|\eta^n(T)|^2\to0.
1191: $$
1192: Let $[x](t)=\sum_{s\in[0,t]}|\Del x_s|^2$.
1193: By the Burkholder-Davies-Gundy inequality (see \cite{delmey}, VII.92),
1194: $$
1195: \bar E|\eta^n(T)|^2 \le c_1\bar E[\eta^n](T),
1196: $$
1197: where $c_1$ is a constant.
1198: Since each jump is bounded by $c_2n^{-1}$
1199: ($c_2$ a constant) and the total number of jumps
1200: $N_n(T)$ satisfies (\ref{ineq:Nn}),
1201: $$
1202: \bar E|\eta^n(T)|^2 \le c_3n^{-2}\bar EN_n(T) \le c_4n^{-1},
1203: $$
1204: which proves (\ref{conv:mgale}).
1205: This implies that $\sup_{[0,T]}|\bar Y^n(t)-\psi^n(t)|\to0$ in
1206: distribution, and therefore the continuity of $\Gamma$ implies $\sup_{[0,T]}|\bar X^n(t)-\phi^n(t)|\to0$
1207: in distribution.
1208: By Lemma \ref{lem:cont}.1,
1209: \beaa
1210: \bar P_{x_n} \left( \hat\sig^n_a \le \bar\sigma^n \right)
1211: &\leq &
1212: \bar P_{x_n}(\bar X^n(\hat\sig_a^n)\in G,
1213: \phi^n(\hat\sig_a^n)\in\pl_{co}G_a)\\
1214: &\le&
1215: \bar P_{x_n} \left (\sup_{t\in[0,T]}|\bar X^n(t)-\phi^n(t)|\ge b_1 \right ),
1216: \eeaa
1217: where $b_1>0$ depends only on $a$.
1218: Hence by (\ref{conv:mgale}),
1219: $\bar P_{x_n}[ \hat\sig^n_a \le \bar\sigma^n]\to0$
1220: as $n\to\infty$. Therefore (\ref{ineq3}) implies
1221: \[
1222: \limsup_{n \goto \infty} V^n(x_n) \leq C_a(x,\beta).
1223: \]
1224: This gives (\ref{eq:ca}) and completes the proof of (\ref{eq:ub}).
1225:
1226:
1227: \noi{\em \uu{Lower bound}}
1228:
1229: \noi
1230: Next we prove
1231: \be\label{eq:lb}
1232: \liminf_{n\to\infty} V^n(x_n) \geq V^+(x).
1233: \ee
1234: By Lemma \ref{lem:cont}.2, it is enough to show that
1235: for all $a<0$
1236: \[
1237: \liminf_{n\to\infty} V^n(x_n) \geq V_a^+(x).
1238: \]
1239: Let $\al\in A$ be given,
1240: and set $C_a(x,\al)=\inf_{m\in\bar M^b} C_a(x,\al[m],m)$.
1241: Then it suffices to show
1242: \be\label{eq:ca2}
1243: \liminf_{n\to\infty} V^n(x_n) \geq C_a(x,\al), \quad a<0.
1244: \ee
1245: Fixing $\al$, we now prove (\ref{eq:ca2}).
1246:
1247: Interchanging the order of infimum and supremum in equation (\ref{eq:dpe})
1248: (see \cite{roc}, Corollary 37.3.2), and noting that
1249: the infimum over $m$ is of a continuous function with
1250: compact level sets, we denote by $m^n(y)$ a point where
1251: the infimum is achieved.
1252: By Lemma~\ref{lem:bound:m}, the components
1253: $ \bar\lambda_i^n (y) $ and $ \bar\mu_i^n (y) $ of $ m^n (y) $
1254: are all bounded by $b_0 $.
1255: For $u\in U$ and $y\in G^n$,
1256: \be\label{ineq12}
1257: 0 \geq \calL^{n,u,m^n(y)}V^n(y) +c+\rho(u,m^n(y)).
1258: \ee
1259: \begin{lemma}\label{lem:xbar2}
1260: Let $n$ be fixed, and let $\al$ and $x_n$ be as above.
1261: Then there is a filtered probability space
1262: $(\bar\Om,\bar F,(\bar F_t),\bar P)$, and $\bar F_t$-adapted
1263: RCLL processes
1264: $\bar X^n$, $\bar Y^n$ and $u^n$ such that with $\bar P$-probability one
1265: $u^n(t)=\al[\bar m^n](t)$ a.e.\ $t$, $\bar m^n(t)=m^n(\bar X^n(t))$,
1266: $\bar X^n=\Gamma(\bar Y^n)$,
1267: $\bar X^n(0)=\bar Y^n(0)=x_n$,
1268: and for any $f$,
1269: \[
1270: f(\bar X^n(t)) - \int_0^t \calL^{n, u^n(s),m^n}f(\bar X^n(s))ds
1271: \]
1272: \[
1273: f(\bar Y^n(t)) - \int_0^t \calL_0^{n, u^n(s), m^n}f(\bar Y^n(s))ds
1274: \]
1275: are $(\bar F_t)$-martingales.
1276: \end{lemma}
1277:
1278: \proof See Section \ref{sec:lemmas}.
1279:
1280:
1281: Let $\bar m^n(t)= m^n(\bar X^n(t))$
1282: and let $\bar\sigma^n$ be the first
1283: exit time of $\bar X^n$ from $G$.
1284: By (\ref{ineq12}) and Lemma \ref{lem:xbar2}, for any bounded stopping
1285: time $S\le\bar\sig^n$,
1286: \be\label{ineq22}
1287: V^n(x_n) \geq \bar E_{x_n} \left[ V^n(\bar X^n(S))
1288: +\int_0^{S}[c+\rho(\al[\bar m^n](s),\bar m^n(s))]ds \right].
1289: \ee
1290: Denoting $\bar m^n(t)=((\bar\la_i^n(t),\bar\mu_i^n(t))$,
1291: define $\phi^n$ as $\phi^n=\Gamma(\psi^n)$, where
1292: \[
1293: \psi^n = x+\int_0^\cdot v(\bar u^n,\bar m^n)ds,
1294: \]
1295: and let
1296: \[
1297: \hat\sig^n_a \doteq \inf\{t:\phi^n (t)\not \in G_a\}.
1298: \]
1299: Then the definition of $C_a(x,\alpha)$ implies
1300: \[
1301: \int_0^{\hat\sig^n_a}[c+\rho(\al[\bar m^n](s),\bar m^n(s))]ds
1302: \geq C_a(x,\al).
1303: \]
1304: Apply (\ref{ineq22}) with $S=\hat\sig_a^n\w\bar\sig^n\w T$,
1305: with large enough $T$,
1306: using the fact that $V^n\ge0$ to get
1307: \beaa
1308: V^n(x_n) &\geq& \bar E_{x_n} \left[
1309: \int_0^{\hat\sig^n_a\w\bar\s^n\w T}
1310: [c+\rho(\al[\bar m^n](s), \bar m^n(s))]ds \right]\\
1311: &\ge&
1312: \bar E_{x_n}\left[1_{\hat\sig^n_a\le\bar\sig^n}
1313: \int_0^{\hat\sig^n_a\w T}
1314: [c+\rho(\al[\bar m^n](s), \bar m^n(s))]ds \right]\\
1315: &\ge&
1316: \bar P_{x_n}(\hat\s^n_a\le\bar\s^n)C_a(x,\al).
1317: \eeaa
1318: The proof that
1319: $\bar P_{x_n}(\hat\sig^n_a \le \bar\sigma^n)$ tends to one
1320: is analogous to the proof of the that
1321: $\bar P_{x_n}(\hat\sig^n_a \le \bar\sigma^n)\to0$ in the
1322: upper bound. It is therefore omitted. Hence
1323: \[
1324: \liminf_{n \goto \infty} V^n(x_n) \geq C_a(x,\al).
1325: \]
1326: This gives (\ref{eq:ca2}),
1327: and the proof of (\ref{eq:lb}) is established.
1328: \qed
1329:
1330: In fact, the value of the game is independent of $b$ for large $b$,
1331: so that the game has a value with the unbounded action space $M$.
1332: As the result depends on Theorems \ref{th:limit}, \ref{th:unique},
1333: \ref{th:solve} we postpone the proof to Section~\ref{sec:lemmas}.
1334: \begin{theorem}\label{th:b}
1335: For all $b\in[b^*,\infty]$, $V^{b,+}=V^+=V^{b,-}=V^-$.
1336: \end{theorem}
1337: \proof See Section \ref{sec:lemmas}.
1338:
1339:
1340: \section{The PDE}\label{sec:pde}
1341:
1342: In this section we show that the upper and lower values of the game
1343: are the unique Lipschitz viscosity solutions of the PDE (\ref{eq:pde}).
1344: Throughout, the parameter $b\in[b^*,\infty)$ is fixed.
1345: Let
1346: \begin{equation}\label{def:H}
1347: H(q)=\inf_m\sup_u[\lan q,v(u,m)\ran+\rho(u,m)+c].
1348: \end{equation}
1349: It will be useful to
1350: note that the infimum is over the compact set $M^b$,
1351: and the map $(q,u,m)\mapsto [\lan q,v(u,m)\ran+\rho(u,m)+c]$
1352: is continuous.
1353: The PDE of interest is
1354: \be\label{eq:pde}
1355: \lt\{
1356: \begin{array}{ll}
1357: H(DV(x))=0, & x\in G^o,\\
1358: \lan DV(x),\gamma_i\ran=0, & i\in I(x),\ x\in\pl_+G ,\\
1359: V(x)=0, & x\in\OBdry G.
1360: \end{array}\rt.
1361: \ee
1362: Here, $\gamma_i$ are the directions of constraint that were introduced in Section 2.
1363: \begin{definition}
1364: Let a Lipschitz continuous function $u:X \to\R$ be given
1365: (where $X\subset G$).
1366: We say that $u$ is a subsolution [respectively, supersolution]
1367: to (\ref{eq:pde}) on $X$ if the following conditions hold. Let
1368: $\theta: X \to\R$ be continuously differentiable on $\bar X$.
1369: Let $y\in X$ be a local maximum [minimum] of the map
1370: $x\mapsto u(x)-\theta(x)$. Then
1371: \be\label{eq:subs}
1372: H(D\theta(y))\vee\max_{i\in I(y)}\lan D\theta(y),\gamma_i\ran\ge0,
1373: \ee
1374: \be\label{eq:supers}
1375: [\quad
1376: H(D\theta(y))\w\min_{i\in I(y)}\lan D\theta(y),\gamma_i\ran\le0,
1377: \quad]
1378: \ee
1379: and
1380: \be\label{eq:subsb}
1381: V(x)\le0,\quad x\in\bar X\cap\OBdry G,
1382: \ee
1383: \be\label{eq:supersb}
1384: [\quad
1385: V(x)\ge0,\quad x\in\bar X\cap\OBdry G.
1386: \quad]
1387: \ee
1388: We say that $V$ is a viscosity solution to (\ref{eq:pde}),
1389: if it is both a subsolution on $G$ and a supersolution
1390: on $G\setminus\pl_cG$.
1391: \end{definition}
1392:
1393: \remark
1394: In case that $\pl_cG\ne\emptyset$, a viscosity solution is often
1395: called a constrained viscosity solution (cf.\ Soner \cite{son},
1396: Capuzzo-Dolcetta and Lions~\cite{capLion}).
1397: The requirement that $V$ is a subsolution up to the boundary
1398: $\pl_cG$---the part of the boundary where exit can be unilaterally
1399: blocked---serves as a boundary condition on this
1400: part of the boundary. Note that in the current paper,
1401: the term `constrained' refers to the part $\pl_+G$ of the
1402: boundary, where it is the mechanism associated with the Skorokhod
1403: Problem that constrains the dynamics to $G$.
1404: \qed
1405:
1406: First, we address uniqueness of solutions to (\ref{eq:pde}).
1407: \begin{theorem}\label{th:unique}
1408: Let $u$ be a subsolution and $v$ a supersolution to (\ref{eq:pde}).
1409: Then $u\le v$ on $G$.
1410: \end{theorem}
1411:
1412: The proof combines ideas from two sources, namely \cite{atadup2} (which is based on \cite{dupish2}, and discusses how to deal with the constrained
1413: dynamics on $\pl_+G$), and \cite{son} (to accommodate the fact that under Condition \ref{cond:G}.1
1414: part of the boundary ($\pl_cG$) can be thought
1415: of as imposing a state-space constraint on the maximizing player).
1416:
1417: The following lemma will be used in proving Theorem \ref{th:unique}.
1418: In the interest of consistency with previous
1419: publications, we use $B$ in Lemma~\ref{lem:mu} below to denote a certain
1420: subset of $\R^J$ (although everywhere except in this section, $B$ denotes
1421: a set of strategies).
1422: Part 1 states that the ``Set B'' Condition holds, namely a condition
1423: under which it was proved in \cite{dupish1} that the SM enjoys
1424: the regularity property (\ref{sp:lip}).
1425: The proof that this condition holds in
1426: the current setting can be found in \cite{dupram23}.
1427: The existence of a smooth version of the set $B$ is proved in
1428: \cite{atadup2}
1429: (before Lemma 2.1). For Parts 2 and 3, see Lemmas 2.1 and 2.2 of
1430: \cite{atadup2} (note that the condition that $\gamma_i$ are independent
1431: holds).
1432: \begin{lemma}\label{lem:mu}
1433: \ben
1434: \item
1435: There exists a compact, convex, and symmetric set $B\subset \RJ $ with
1436: $0\in B^o$, such that if $z\in\pl B$ and if $n$ is an outward
1437: normal to $B$ at $z$, then for all $i\in\{1,\ldots,J\}$
1438: $$
1439: \lan z,e_i\ran\ge-1 {\rm\ implies\ } \lan\gamma_i,n\ran\ge0
1440: {\rm\ and\ } \lan z,e_i\ran\le1 {\rm\ implies\ } \lan\gamma_i,n\ran\le0.
1441: $$
1442: In addition, the unit outward normal $n(x)$ to $B$ at $x\in\pl B$ is
1443: unique and continuous (as a function on $\pl B$).
1444: \item
1445: Let $\bar n$ be the extension of $n$ to $ \RJ $
1446: satisfying $\bar n(x)=n(y)$ whenever $ax=y\in\pl B$, some $a\in(0,\infty)$
1447: (and define $\bar n(0)$ arbitrarily).
1448: Let $\Xi: \RJ \to\R_+$ be defined via
1449: $$
1450: \Xi(x)=a\ \Leftrightarrow\ x\in\pl(aB)
1451: $$
1452: for all $a\in[0,\infty)$, and let $\xi(x)=(\Xi(x))^2$.
1453: Then there exist constants $m,M\in(0,\infty)$ and a function
1454: $\vr: \RJ \to[m,M]$ such that the $C^1( \RJ )$
1455: function $\xi$ satisfies $m\|x\|^2\le\xi(x)\le M\|x\|^2$, and
1456: $D\xi(x)=\vr(x)\Xi(x)\bar n(x)$.
1457: \item
1458: There exists a constant $m_1\in (0,\infty)$ and a continuously differentiable
1459: function $\mu: \RJP \to[0,m_1]$ such that $\|D\mu\|\le m_1$ on
1460: $ \RJP $, and
1461: $$
1462: \lan D\mu(x),\gamma_i\ran<0,\quad x\in \RJP ,\quad i\in I(x).
1463: $$
1464: \een
1465: \end{lemma}
1466:
1467:
1468: In what follows, we keep the notation of Lemma \ref{lem:mu}
1469: for $B, \bar n, \Xi, \xi, \vr$ and $\mu$.
1470:
1471:
1472: \noindent{\bf Proof of Theorem \ref{th:unique}:}
1473: For $a>0$, let
1474: $$
1475: U(x)=u(x)-a\mu(x),
1476: $$
1477: $$
1478: V(x)=v(x)+a\mu(x).
1479: $$
1480: Let $\del>0$. Then it suffices to show that for all small $a>0$,
1481: $\del>0$, one has $U\le (1+\del)V$ on $G$. Arguing by contradiction,
1482: we assume that this is not true. Then there are $a$ and $\del$
1483: arbitrarily small such that
1484: $$
1485: \rho=\sup_{x\in G}[U(x)-(1+\del)V(x)]>0.
1486: $$
1487: Below we let $c_i, i=1,2,...$ denote positive constants.
1488: Consider Condition \ref{cond:G}.1 first.
1489: Let
1490: \begin{equation}\label{eq:phixy}
1491: \Phi(x,y)= U(x)-(1+\del)V(y)-\frac 1 \eps\xi(x -y-\eps^{1/2}y).
1492: \end{equation}
1493: Let $(\bar x,\bar y)\in \bar G^2$
1494: achieve the maximum of $\Phi$ in $ \bar G \times
1495: \bar G $.
1496: By continuity of $U$ and $V$, there exists $ \bar z\in\bar G $ so that
1497: $ \rho = U( \bar z)-(1+\del)V(\bar z) $. Note that
1498: \begin{equation}\label{eq:10}
1499: (1+\eps^{1/2})^{-1}\bar z\in\bar G.
1500: \end{equation}
1501: Hence by the Lipschitz continuity
1502: of $V$,
1503: \bea\label{e:phi1}\nonumber
1504: \Phi ( \bar x , \bar y )
1505: & \geq & \Phi
1506: \left ( \bar z, \frac {\bar z} {1 + \eps^{\Sfrac 1 2 }} \right )
1507: \\
1508: \nonumber
1509: & = & U(\bar z)-(1+\del)
1510: V \left (\frac {\bar z} {1 + \eps^{\Sfrac 1 2 }} \right )
1511: \\
1512: & \geq & \rho - c_1 \eps^{\Sfrac 1 2 } .
1513: \eea
1514: By Lipschitz continuity of $U$ and the lower bound on $\xi$ given in Lemma \ref{lem:mu},
1515: \bea\label{e:phi2}\nonumber
1516: \Phi ( \bar x , \bar y )
1517: & = & U(\bar x)-(1+\del)V(\bar y)
1518: -\frac 1 \eps \xi (\bar x - \bar y-\eps^{1/2}\bar y) \\
1519: \nonumber
1520: & \leq & U (\bar y ) + c_2 | \bar x - \bar y | - (1+\del)V(\bar y)
1521: - \frac m \eps | \bar x - \bar y-\eps^{1/2}\bar y |^2\\
1522: &\le&
1523: \rho + c_2 | \bar x - \bar y |
1524: - \frac{c_3}{\eps} | \bar x- \bar y-\eps^{1/2}\bar y |^2 .
1525: \eea
1526: By (\ref{e:phi1}) and (\ref{e:phi2}),
1527: \bea\label{e:phi3}
1528: c_2 | \bar x - \bar y | + c_1 \eps^{\Sfrac 1 2}
1529: &\geq&
1530: \frac{c_3}{\eps} | \bar x- \bar y-\eps^{1/2}\bar y |^2
1531: \\ \nonumber
1532: &\geq&
1533: \frac{c_4}{\eps} | \bar x - \bar y |^2 - c_4 | \bar y |^2 .
1534: \eea
1535: Since $ \bar x $ and $ \bar y $ are bounded,
1536: (\ref{e:phi3}) implies
1537: $| \bar x - \bar y |^2 \leq c_5 \eps$
1538: and so
1539: \begin{equation}\label{e:phi5}
1540: | \bar x - \bar y | \leq c_6 \eps^{\Sfrac 1 2 }.
1541: \end{equation}
1542: Using this in~(\ref{e:phi3}) we have
1543: \begin{equation}\label{e:phi4}
1544: | \bar x- \bar y-\eps^{1/2}\bar y | \leq c_7 \eps^{\Sfrac 3 4} .
1545: \end{equation}
1546: By (\ref{e:phi5}),
1547: $\bar x-\bar y\to0$ as $\eps\downarrow0$. Also,
1548: we claim that
1549: for all $\eps>0$ small, $\bar x$ and $\bar y$ are
1550: bounded away from $\pl_o G$.
1551: To see this, assume the contrary. Then along a subsequence, both
1552: $\bar x$ and $\bar y$ must converge to the same point on $\pl_o G$.
1553: Using the continuity of $u$ and $v$,
1554: (\ref{eq:subsb})--(\ref{eq:supersb}),
1555: and the non-negativity of $\xi$,
1556: $\limsup\Phi(\bar x,\bar y)
1557: \le \limsup \left [ u(\bar x)-(1+\del)v(\bar y) \right ] \le0$,
1558: where the limit superior is taken along this subsequence.
1559: However, by (\ref{e:phi1}),
1560: for all small $\eps$, $\Phi(\bar x,\bar y)\ge\rho/2>0$, which
1561: gives a contradiction.
1562:
1563: Let
1564: $$
1565: \theta(x)=\frac1\eps\xi(x-\bar y-\eps^{1/2}\bar y)+a\mu(x),
1566: $$
1567: and note that the map $x\mapsto u(x)-\theta(x)$ has a maximum
1568: at $\bar x\in G$.
1569: Since $u$ is a subsolution, (\ref{eq:subs}) must
1570: be satisfied at $\bar x$. Denoting
1571: \begin{equation}\label{eq:que}
1572: q^\eps=
1573: \vr(\bar x-\bar y-\eps^{1/2}\bar y)\Xi(\bar x-\bar y-\eps^{1/2}\bar y)
1574: \bar n(\bar x-\bar y-\eps^{1/2}\bar y),
1575: \end{equation}
1576: we have from Lemma \ref{lem:mu}.2 that
1577: $$
1578: D\theta(\bar x)=\frac1\eps q^\eps+aD\mu(\bar x).
1579: $$
1580: Suppose $i$ is such that $\bar x_i=0$. Then
1581: $ \lan\bar x-\bar y-\eps^{1/2}\bar y,e_i\ran\le0 $
1582: and so by Lemma \ref{lem:mu}.1,
1583: $$
1584: \lan\gamma_i, \bar n(\bar x-\bar y-\eps^{1/2}\bar y)\ran\le0.
1585: $$
1586: Since by Lemma \ref{lem:mu}.3, $\lan\gamma_i, D\mu(\bar x)\ran<0$,
1587: it follows that $\lan\gamma_i,D\theta(\bar x)\ran<0$. It follows from
1588: (\ref{eq:subs}) that
1589: $$
1590: H(D\theta(\bar x))\ge0,
1591: $$
1592: namely,
1593: \be\label{eq:H1}
1594: H\lt(\frac1\eps q^\eps+aD\mu(\bar x)\rt)\ge0.
1595: \ee
1596: On the other hand, let
1597: $$
1598: \al(y)=-a\mu(y)-\frac{1}{\eps(1+\del)}\xi(\bar x-y-\eps^{1/2}y).
1599: $$
1600: Note that
1601: $$
1602: D\al(\bar y)=-aD\mu(\bar y)+\frac{1+\eps^{1/2}}{\eps(1+\del)} q^\eps
1603: $$
1604: and that
1605: the map $y\mapsto v(y)-\al(y)$ has a minimum at $\bar y$.
1606: Since $\bar x\in\bar G$, (\ref{e:phi4}) implies that
1607: $\bar y\in G\setminus\pl_cG$ for all small $\eps$.
1608: Since $v$ is a supersolution,
1609: (\ref{eq:supers}) is satisfied at $y$. An argument as above shows
1610: that
1611: $$
1612: H(D\al(\bar y))\le0,
1613: $$
1614: and therefore
1615: $$
1616: H\lt(\frac{1}{1+\del}\lt(\frac{1+\eps^{1/2}}\eps q^\eps
1617: -a(1+\del)D\mu(\bar y)\rt)\rt)\le0.
1618: $$
1619: It follows from the definition of $H$, using $\rho(u,m)\ge0$, that
1620: $$
1621: H\lt(\frac 1{1+\del}\, p\rt)\ge\frac 1{1+\del}H(p)+\frac\del{1+\del}c,
1622: $$
1623: and therefore
1624: \be\label{eq:H2}
1625: H\lt(\frac1\eps q^\eps+\frac1{\eps^{1/2}}q^\eps
1626: -a(1+\del)D\mu(\bar y)\rt)+\del c\le0.
1627: \ee
1628: Now $D\mu$ is bounded, and by (\ref{e:phi4}), boundedness of
1629: $n$ and $\vr$, and
1630: the Lipschitz continuity of $\Xi$, it follows that
1631: $\eps^{-1/2}q^\eps$ converges to zero as $\eps\to0$.
1632: Note that by (\ref{def:H}) and the following comment,
1633: $H$ is uniformly continuous on $\R^J$.
1634: Therefore,
1635: (\ref{eq:H1}) and (\ref{eq:H2}) give a contradiction when $a>0$
1636: and $\eps>0$ are small and $\del>0$ fixed.
1637:
1638: Under Condition \ref{cond:G}.2 the above argument
1639: is not valid, since (\ref{eq:10}) may not hold.
1640: However,
1641: in this case the minimizing player can force exit from any point on $\partial G\backslash \partial_+ G$,
1642: and the additional complications due to the ``state-space constraint'' used under part 1 are no longer needed.
1643: In other words,
1644: instead of (\ref{eq:phixy}) we can consider
1645: $$
1646: \Phi(x,y)=U(x)-(1+\del)V(y)-\frac1\eps\xi(x-y),
1647: $$
1648: and a review of the above proof shows that (\ref{eq:H1})
1649: and (\ref{eq:H2}) still hold if the expression
1650: $\eps^{1/2}$ is replaced by zero everywhere in (\ref{eq:que}) and
1651: (\ref{eq:H2}). A contradiction is then obtained analogously.
1652: \qed
1653:
1654: We next consider the upper and lower values of the game,
1655: and remind the reader that in this section the rates $m$ are assumed bounded.
1656: \begin{theorem}\label{th:solve}
1657: $V^-$ and $V^+$ are solutions to (\ref{eq:pde}).
1658: \end{theorem}
1659: Recall that from Lemma \ref{lem:vallip}, $V^\pm$ are Lipschitz.
1660:
1661:
1662:
1663: \noindent
1664: {\bf Proof of Theorem \ref{th:solve}:}
1665: We use the specific form of $v(u,m)$ and $\rho(u,m)$. These can be
1666: written as
1667: $$
1668: v(u,m)=b_0(m)+\sum_{i=1}^Ju_ib_i(m),
1669: $$
1670: $$
1671: \rho(u,m)=c_0(m)+\sum_{i=1}^Ju_ic_i(m).
1672: $$
1673: We have $\sum_{i\in C(k)}u_i\le1$.
1674: The $b_i$ are linear, and the $c_i$ are convex.
1675: Hence, as a direct consequence of \cite[Corollary 37.3.2]{roc},
1676: the {\em Isaacs condition} holds, namely
1677: \be\label{eq:isaacs}
1678: H(q)=\inf_m\sup_u[\lan q,v(u,m)\ran+c+\rho(u,m)]
1679: =\sup_u\inf_m[\lan q,v(u,m)\ran+c+\rho(u,m)].
1680: \ee
1681: Another fact that we will use is that for any
1682: $y\in G\setminus\pl_cG$ there
1683: is $\del_0=\del_0(y)>0$ which serves as a lower bound on the exit time.
1684: Namely, if $\phi$ solves $\dot\phi=\pi(\phi,v(u,m))$, $\phi(0)=y$,
1685: then
1686: \be\label{eq:boundsig}
1687: \sig \doteq \inf\{ t\geq 0: \phi(t) \not \in G\} \ge\del_0,\quad u\in \bar U, m\in \bar M.
1688: \ee
1689: The bound is an immediate consequence of the $u$ and $m$ being
1690: uniformly bounded.
1691:
1692: By definition, $ V ^\pm (x) = 0 $ for $ x \in \pl_o G $.
1693: Thus we only need to
1694: establish (\ref{eq:subs})--(\ref{eq:supers}).
1695: The proof consists of four parts.
1696:
1697: \noindent{\em \underline{Proof that $V^-$ is a supersolution on $G \backslash \partial_cG$.}}
1698:
1699: Standard dynamic programming arguments show that for $\del>0$,
1700: \be\label{eq:dp1}
1701: V^-(x)=\inf_\beta\sup_u\lt[\int_0^{\s\w\del}(c+\rho(u,\beta[u]))dt
1702: +V^-(\phi(\s\w\del))\rt],
1703: \ee
1704: where $\phi$ is the solution to $\dot\phi=\pi(\phi,v(u,\beta[u]))$,
1705: with $\phi(0)=x$.
1706: Let $\theta$ be smooth, and let $y\in G\setminus\pl_cG$
1707: be a local minimum of
1708: $V^--\theta$. We can assume without loss that $V^-(y)=\theta(y)$.
1709: We need to show
1710: \be\label{alpha}
1711: H(D\theta(y))\w\min_{i\in I(y)}\lan D\theta(y),\gamma_i\ran\le0.
1712: \ee
1713: We shall assume the contrary and reach a contradiction.
1714: Thus, there exists $a>0$ such that
1715: $H(D\theta(y))\ge a$, and
1716: \be\label{eq:D1}
1717: \lan D\theta(y),\gamma_i\ran\ge a,\quad i\in I(y).
1718: \ee
1719: From the definition of $H$ and (\ref{eq:isaacs}),
1720: $$
1721: \sup_u\inf_m[\lan D\theta(y),v(u,m)\ran+c+\rho(u,m)]\ge a,
1722: $$
1723: and therefore there exists a $u_0$ such that for all $m$,
1724: $$
1725: \lan D\theta(y),v(u_0,m)\ran+c+\rho(u_0,m)\ge a/2.
1726: $$
1727: For any strategy $\beta$, if $\bar u(t)\equiv u_0$,
1728: \be\label{eq:D2}
1729: \lan D\theta(y),v(\bar u(t),\beta[\bar u](t))\ran+c
1730: +\rho(\bar u(t),\beta[\bar u](t))\ge a/2
1731: \ee
1732: for all $t$. Let $\phi$ denote the dynamics corresponding to $\bar u$
1733: and a generic $\beta$, starting from $y$.
1734: Note that the mapping $z\mapsto I(z)$ is upper semi-continuous, in
1735: the sense that for any $z$ there is a neighborhood of $z$ on which
1736: $I(\cdot)\subset I(z)$. Using the boundedness of $m$
1737: this implies that for any $\beta\in B$, one has
1738: $I(\phi(r))\subset I(y)$ for $r\in[0,\del]$, if $\del>0$ is
1739: chosen small enough.
1740: We now use that $\phi$ is a solution to the SP.
1741: Choosing such a $\del >0$, for any $r\in[0,\del]$ there
1742: exist $a_i\ge0$ (that may depend on $r$) such that
1743: $$
1744: \dot\phi(r)=v(\bar u(r),\beta[\bar u](r))+\sum_{i\in I(y)}a_i\gamma_i.
1745: $$
1746: Using the continuity of $D\theta$ and taking $\del>0$ smaller
1747: if necessary, (\ref{eq:D1}) and (\ref{eq:D2}) imply,
1748: for $t\in[0,\del]$,
1749: \beaa
1750: \frac{d}{dt}\theta(\phi(t)) &=&
1751: \lan D\theta(\phi(t)),\dot\phi(t)\ran\\
1752: &\ge&
1753: -c-\rho(\bar u(t),\beta[\bar u](t))+a/4.
1754: \eeaa
1755: Taking $\del$ even smaller if necessary (so that it is at most
1756: $\del_0$), we have from (\ref{eq:boundsig}) that
1757: $$
1758: \theta(\phi(\del))-\theta(y)\ge
1759: -\int_0^{\del}(c+\rho(\bar u(t),\beta[\bar u](t)))dt
1760: +a\del/4.
1761: $$
1762: From (\ref{eq:dp1}), one can find a $\beta$ such that
1763: $$
1764: V^-(y)\ge\sup_u\lt[\int_0^{\del}(c+\rho(u,\beta[u]))dt
1765: +V^-(\phi(\del))-a\del/8\rt].
1766: $$
1767: Letting $u=\bar u$, the last two displays give (using $\theta(y)
1768: =V^-(y)$)
1769: $$
1770: \theta(\phi(\del))\ge V^-(\phi(\del))+a\del/8,
1771: $$
1772: so that $V^-(\phi(\del))-\theta(\phi(\del))<0$ for all
1773: $\del>0$ small, contradicting the assumption that $y$ is
1774: a local minimum of $V^--\theta$. This proves that $V^-$ is a
1775: supersolution on $G \backslash \partial_cG$.
1776:
1777: \noindent
1778: {\em \underline{Proof that $V^-$ is a subsolution on $G$.}}
1779:
1780: Let $\theta$ be smooth and $y\in G$ a local maximum of
1781: $V^--\theta$.
1782: In case that $y\in\pl_cG$, let $\bar U_{y,\beta,\del}$
1783: be the set
1784: of controls $u\in\bar U$ for which the trajectory $\phi$ determined
1785: by $u$ and $\beta[u]$ and starting from $y$ does not exit $G$
1786: on $[0,\del]$. Given $y\in\pl_cG$, it is clear that
1787: $\bar U_{y,\beta,\del}$ is not empty for all $\del$
1788: small and all $\beta$, by considering the control $u=0$.
1789: Moreover, for all $\del$ small enough,
1790: (\ref{eq:dp1}) is valid where the supremum
1791: extends only over $u\in\bar U_{y,\beta,\del}$. Indeed, given
1792: $u\not\in\bar U_{y,\beta,\del}$, consider
1793: $u'$ that agrees with $u$ on $[0,\sig]$ and $u'=0$ on
1794: $(\sig,\del]$. Then the expression in brackets in (\ref{eq:dp1})
1795: is identical under $u$ and under $u'$, but
1796: $u'\in\bar U_{y,\beta,\del}$.
1797:
1798: Assume without loss that $V^-(y)=\theta(y)$.
1799: We would like to show that
1800: \be\label{gamma}
1801: H(D\theta(y))\vee\max_{i\in I(y)}\lan D\theta(y),\gamma_i\ran\ge0.
1802: \ee
1803: Assuming the contrary, there exists $a>0$ such that
1804: $H(D\theta(y))\le-a$, and
1805: \be\label{eq:D3}
1806: \lan D\theta(y),\gamma_i\ran\le-a,\quad i\in I(y).
1807: \ee
1808: Using the definition of $H$ and (\ref{eq:isaacs}), for all $u$ there exists $m_u$ such that
1809: \be\label{eq:D4}
1810: \lan D\theta(y),v(u,m_u)\ran
1811: +c+\rho(u,m_u)\le-a/2.
1812: \ee
1813: Note that it is possible to choose $m_u$ so that it depends continuously
1814: on $u$.
1815: Define $\bar\beta$ as $\bar\beta[u](t)=m_{u(t)}$ for all $t$. Since
1816: $\bar\beta[u]$ is measurable if $u$ is, $\bar\beta$ maps $\bar U$
1817: into $\bar M$.
1818: Let $\phi$ be the trajectory corresponding to $\bar\beta$ and a generic
1819: $u\in \bar U$, (or a generic $u\in\bar U_{y,\beta,\del}$ if
1820: $y\in\pl_cG$) starting from $y$.
1821: Arguing as before by upper semi-continuity
1822: of $I(\cdot)$, if $\del$ is small enough, then
1823: $$
1824: \dot\phi(r)=v(u(r),\bar\beta[u](r))+\sum_{i\in I(y)}a_i\gamma_i,
1825: \quad r\in[0,\del],
1826: $$
1827: where $a_i\ge0$ may depend on $r$.
1828: By possibly taking $\del$ smaller, and smaller
1829: than $\del_0$, we have, using the continuity of $D\theta$ and
1830: (\ref{eq:D3}), (\ref{eq:D4}) that
1831: \beaa
1832: \frac{d}{dt}\theta(\phi(t)) &=&
1833: \lan D\theta(\phi(t)),\dot\phi(t)\ran\\
1834: &\le&
1835: -c-\rho(u(t),\bar\beta[u](t))-a/4,
1836: \eeaa
1837: and
1838: $$
1839: \theta(\phi(\del))-\theta(y)\le-\int_0^\del(c+
1840: \rho(u(t),\bar\beta[u](t)))dt-a\del/4.
1841: $$
1842: Now, (\ref{eq:dp1}) implies that for any $\beta$ there
1843: is $u$ such that
1844: $$
1845: V^-(y)\le\int_0^\del(c+\rho(u,\beta[u]))dt+V^-(\phi(\del))+a\del/8.
1846: $$
1847: Specializing to $\beta=\bar\beta$, the last two displays show
1848: that $V^-(\phi(\del))-\theta(\phi(\del))>0$ for all $\del>0$
1849: small. This contradicts the assumption that $y$ is a local
1850: maximum of $V^--\theta$, and as a result, $V^-$ is a subsolution.
1851:
1852: \noindent
1853: {\em \underline{Proof that $V^+$ is a supersolution on $G \backslash \partial_cG$.}}
1854:
1855: The proof is analogous to the proof that $V^-$ is a subsolution.
1856: Most details are therefore skipped.
1857: The dynamic programming principle states that
1858: for $\del>0$,
1859: \be\label{eq:dp2}
1860: V^+(x)=\sup_\al\inf_m\lt[\int_0^{\s\w\del}(c+\rho(\al[m],m))dt
1861: +V^+(\phi(\s\w\del))\rt],
1862: \ee
1863: where $\phi$ is the dynamics corresponding to $\al$ and $m$,
1864: starting from $x$. Taking a smooth $\theta$, and leting
1865: $y\in G\setminus\pl_cG$
1866: be a local minimum of $V^+-\theta$, showing
1867: $$
1868: H(D\theta(y))\w\min_{i\in I(y)}\lan D\theta(y),\gamma_i\ran\le0
1869: $$
1870: can be obtained by an argument analogous to that used to prove (\ref{gamma}),
1871: using (\ref{eq:dp2}) in place of (\ref{eq:dp1}).
1872:
1873: \noindent
1874: {\em \underline{Proof that $V^+$ is a subsolution on $G$.}}
1875:
1876: We need to show that
1877: \be\label{beta}
1878: H(D\theta(y))\vee\max_{i\in I(y)}\lan D\theta(y),\gamma_i\ran\ge0,
1879: \ee
1880: where $\theta$ is smooth, and $y\in G$ is a local maximum of
1881: $V^+-\theta$.
1882: In the special case where $y\in\pl_cG$, we can assume
1883: without loss that the supremum in (\ref{eq:dp2}) extends
1884: only over $\al\in A_{y,\del}$, the set of strategies
1885: under which, for any $m\in\bar M^b$, the dynamics associated
1886: with $\al$ and $m$, and starting from $y$,
1887: does not leave $G$ before $\del$.
1888: The proof of (\ref{beta})
1889: is analogous to the proof of (\ref{alpha}), and is skipped.
1890:
1891:
1892: This completes the proof that $V^-$ and $V^+$ are solutions
1893: to (\ref{eq:pde}).
1894: \qed
1895:
1896:
1897:
1898: \section{A competing queues example}\label{sec:example}
1899:
1900: Consider a queueing network with only one server, providing
1901: service to $J$ classes. Each customer requires service once.
1902: In this example all arrival rates are positive:
1903: $\la_i>0$ for all $i$, hence $\calJ_+=\{1,\ldots,J\}$.
1904: This network, ``the $k$ competing queues,'' has been studied extensively,
1905: in discrete and continuous
1906: time (see~\cite{BDM,W} and references therein). When the criterion
1907: (to be minimized) is
1908: either the average cost or the discounted cost, and the one-step cost
1909: is a positive linear combination
1910: $\sum_i c_ix_i$ of the queue sizes $x_i$, the optimal policy is
1911: the $ \mu $-$c$ rule, which is a priority discipline,
1912: giving absolute priority to the non-empty queue for which
1913: $\mu_ic_i$ is maximal.
1914: Under the cost studied here, the optimal policy is quite different.
1915:
1916: \begin{proposition}
1917: Consider the case where $G$ is a hyper-rectangle, given
1918: as $G=\{x:0\le x<z_i\}$, where $z_i>0$ are constants.
1919: Assume that $\la_i>0$ for all $i=1,\ldots,J$.
1920: If $c$ is large enough,
1921: then the viscosity solution to the PDE (\ref{eq:pde}) is given as
1922: \begin{equation}\label{eq:form}
1923: V(x)=\min_i\al_i(z_i-x_i),
1924: \end{equation}
1925: where $\al_i>0$ are constants depending on $c$.
1926: \end{proposition}
1927: We remark that the constants $\al_i$ are uniquely defined by
1928: (\ref{eq:form1}) below.
1929: In the totally symmetric case, where $\mu_i=\mu$, $\la_i=\la$,
1930: $z_i=z$ for all $i$, the solution takes the form
1931: $V(x)=\al\min_i(z-x_i)$. In this case, the optimal
1932: service discipline can be interpreted as ``serve the longest
1933: queue.'' An asymmetric two dimensional example
1934: is given in Figure \ref{fig:example}, where the domain $G$
1935: is divided into two subdomains $G_1$ and $G_2$ in accordance
1936: with the structure (\ref{eq:form}), and the optimal service
1937: discipline corresponds to giving priority to
1938: class $i$ when the state is within $G_i$, $i=1,2$.
1939: Thus the optimal control under our escape-time criterion is very different
1940: from the optimal controls for the average or discounted cost criteria.
1941:
1942:
1943: \begin{figure} % export: 30 percent.
1944: \centerline{
1945: \begin{tabular}{cc}
1946: \psfig{file=fig_net_2.eps}\qquad\qquad\qquad
1947: \psfig{file=fig_ex_2.eps}
1948: \end{tabular}
1949: }
1950: \caption{Priority to class $i$ when the state is in $G_i$, $i=1,2$.}
1951: \label{fig:example}
1952: \end{figure}
1953:
1954: \proof
1955: The constraint directions are given by $\gamma_i=e_i$.
1956: The Hamiltonian is given by
1957: $$
1958: H(p)=\sup_u\inf_m H(p,u,m),
1959: $$
1960: where
1961: $$
1962: H(p,u,m)=c+\sum_i\lt[p_i(\bar\la_i-u_i\bar\mu_i)
1963: +\la_i l\left (\frac{\bar\la_i}{\la_i}\right)
1964: +u_i\mu_il\left(\frac{\bar\mu_i}{\mu_i}\right)\rt].
1965: $$
1966: Using strict convexity and smoothness of the map
1967: $m\mapsto H(p,u,m)$, the minimum over $m$ is attained
1968: at $\bar\la_i=\la_i e^{-p_i}$, $\bar\mu_i=\mu_ie^{p_i}$.
1969: Thus
1970: $$
1971: H(p,u)\doteq \inf_m H(p,u,m)=
1972: c+\sum_i[\la_i(1-e^{-p_i})+u_i\mu_i(1-e^{p_i})].
1973: $$
1974: For the proposed solution, $DV(x)\in-\R_+^J$
1975: wherever the gradient is defined. For $p\in-\R_+^J$, maximizing
1976: $H(p,u)$ over $u$ clearly gives
1977: \begin{equation}\label{eq:ham}
1978: H(p)=\sup_u H(p,u)=c+\sum_i\la_i(1-e^{-p_i})+\max_i\mu_i(1-e^{p_i}).
1979: \end{equation}
1980: We use the well known fact that the definition
1981: of viscosity solutions can be equivalently stated in terms
1982: of sub- and superdifferentials (see \cite{barcap}, Lemma II.1.7).
1983: Note that (\ref{eq:subsb}) and (\ref{eq:supersb})
1984: hold, since $V=0$ on $\pl_oG$.
1985: Hence it suffices to verify
1986: that (\ref{eq:subs}) [resp., (\ref{eq:supers})] holds where
1987: $D\theta(y)$ is replaced by any superdifferential [subdifferential]
1988: of $V$ at $y$.
1989:
1990: We show first that the equation $H(DV(x))=0$ holds wherever
1991: $DV$ is defined.
1992: The proposed form (\ref{eq:form}) satisfies $DV(x)=-\al_ie_i$,
1993: wherever the gradient is defined, with $i=i_x$ depending on $x$.
1994: By the special form of the gradient,
1995: the equation $H(DV(x))=0$ takes the form
1996: \begin{equation}\label{eq:form1}
1997: H(DV(x))=c+\la_i(1-e^{\al_i})+\mu_i(1-e^{-\al_i})=0,
1998: \end{equation}
1999: where $i=i_x$. Denote $c_i=c/(\la_i+\mu_i)$.
2000: Then equivalently,
2001: $
2002: 1+c_i-F_i(\al_i)=0,
2003: $
2004: where
2005: $$
2006: F_i(\al_i)=\frac{\la_ie^{\al_i}+\mu_ie^{-\al_i}}{\la_i+\mu_i}.
2007: $$
2008: The function $F_i$ is strictly convex, $F_i(0)=1$,
2009: and $F_i(\al_i)\to\infty$ as $\al_i\to\infty$. Since $c_i>0$,
2010: it follows that there are unique positive constants
2011: $\al_i$ where $F_i(\al_i)=1+c_i$, $i=1,\ldots,J$.
2012: These are the constants
2013: in (\ref{eq:form}). In particular, (\ref{eq:form1}) holds
2014: for $i=i_x$, and $H(DV(x))=0$.
2015:
2016: Next consider any interior point $x$ at which the
2017: gradient is not defined. Clearly
2018: there are no subdifferentials at that point, and any
2019: superdifferential is given as a convex combination of
2020: $-\al_ie_i$, $i=1,\ldots,J$.
2021: Let $B(e_i,\eps)$ be the open ball of radius $\eps$ about $e_i$.
2022: Denote
2023: $\til S=\{\nu\in\R^J:\nu_i\ge0,\sum\nu_i=1\}$,
2024: $S=\{\nu\in\R^J:\nu_i\ge0,\sum\nu_i\le1\}$,
2025: $S_\eps=S\cap\cup_iB(e_i,\eps)$, and $S_\eps^c=S-S_\eps$.
2026: Let $q=-\sum_i\nu_i\al_ie_i$. It suffices to show
2027: that $H(q)\ge0$ for $\nu\in\til S$, but since we later need a stronger
2028: statement than that, we show that in fact $H(q)\ge0$
2029: holds for $\nu\in S$. By (\ref{eq:ham}),
2030: $$
2031: H(q)=c+\sum_i\la_i(1-e^{\nu_i\al_i})+\max_i\mu_i(1-e^{-\nu_i\al_i}).
2032: $$
2033: Define
2034: $$
2035: H^1(q)=c+\sum_i\la_i(1-e^{\nu_i\al_i})+\mu_1(1-e^{-\nu_1\al_1}),
2036: $$
2037: $$
2038: \bar H(q)=c+\sum_i[\la_i(1-e^{\nu_i\al_i})+\mu_i(1-e^{-\nu_i\al_i})].
2039: $$
2040: By (\ref{eq:form1}), $c+\la_i(1-e^{\al_i})+\mu_i\ge0$ and
2041: $c+\la_i(1-e^{\al_i})\le0$, and it follows that there are
2042: constants $A_1,A_2,A_3$ and $A_4$ such that for all $c$ and $i=1,\ldots,J$,
2043: \begin{equation}\label{eq:A}
2044: A_1+\log(c+A_2)\le\al_i\le A_3+\log(c+A_4).
2045: \end{equation}
2046:
2047: We first consider small perturbations $\nu$ of $e_1$.
2048: To show that $H(q)\ge0$, it suffices to show that $H^1(q)\ge0$.
2049: Note that (\ref{eq:form1}) implies $H^1(q)|_{\nu=e_1}=0$. Also,
2050: $$
2051: \nabla_\nu H^1(q)|_{\nu=e_1}
2052: =(-\la_1\al_1e^{\al_1}+\mu_1\al_1e^{-\al_1})e_1
2053: -\sum_{i\ne1}\la_i\al_ie_i.
2054: $$
2055: Hence, for $\gamma=e_i-e_1$
2056: (where $i\ne1$), using $c+\la_1(1-e^{\al_1})\le0$,
2057: (\ref{eq:form1}) and (\ref{eq:A}),
2058: \beaa
2059: \nabla_\nu H^1(q)|_{\nu=e_1}\cdot\gamma
2060: &=&
2061: \la_1\al_1e^{\al_1}-\mu_1\al_1e^{-\al_1}-\al_i\la_i\\
2062: &=&
2063: \al_1(2\la_1e^{\al_1}-\mu_1-c-\la_1)-\al_i\la_i\\
2064: &\ge&
2065: \al_1(c-\mu_1)-\al_i\la_i\\
2066: &\ge&
2067: [A_1+\log(c+A_2)](c-\mu_1)-[A_3+\log(c+A_4)]\la_i\\
2068: &\ge&
2069: 1,
2070: \eeaa
2071: for all $c$ large.
2072: Analogous calculations give $\nabla_\nu H^1(q)|_{\nu=e_1}
2073: \cdot\gamma\ge1$ for $\gamma=-e_1$ as well.
2074: As a result, the directional derivatives
2075: $(\pl/\pl\til\gamma)H^1(q)|_{\nu=e_1}$ in the direction $\til\gamma$,
2076: where $\til\gamma$ are of the form
2077: $\til\gamma=(y-e_1)/\|y-e_1\|$, $y\in S$, are bounded below
2078: by $1/2$. Hence $H^1(q)\ge0$ for $\nu\in S$ within
2079: a neighborhood of $e_1$
2080: and $c$ large. Consequently, a similar statement holds for
2081: $H(q)$. Since the same argument holds for neighborhoods of
2082: $e_i$, $i=2,\ldots,J$, we conclude that there is $\eps>0$
2083: and $c_0$ such that $H(q)\ge0$ for $\nu\in S_\eps$ and $c\ge c_0$.
2084:
2085: Next consider $\nu\in S_\eps^c$.
2086: We first provide a lower bound on $(\pl/\pl c)\bar H(q)$.
2087: Differentiating (\ref{eq:form1}) with respect to $c$,
2088: $\dot\al_i\doteq
2089: \pl\al_i/\pl c=(\la_i e^{\al_i}-\mu_ie^{-\al_i})^{-1}$.
2090: Using (\ref{eq:A}), for all $c$ large,
2091: $0\le\dot\al_i\le(\la_ie^{\al_i}-1)^{-1}$. Using this,
2092: the fact that $\nu$ is bounded away from $\cup_i\{e_i\}$,
2093: and by taking $c$ large, one has
2094: \beaa
2095: \frac{\pl}{\pl c}\bar H(q)
2096: &\ge&
2097: 1-\sum_i\nu_i\la_i\dot\al_ie^{\nu_i\al_i}\\
2098: &\ge&
2099: 1-\sum_i\nu_i[e^{(1-\nu_i)\al_i}-1]^{-1}\\
2100: &\ge&
2101: \frac12.
2102: \eeaa
2103: Note that the above bound holds for all $c\ge c_1$ and
2104: all $\nu\in S_\eps^c$, where $c_1$ is a constant.
2105: It follows that there is $c_2$ such that for all $c\ge c_2$
2106: and all $\nu\in S_\eps^c$, one has $\bar H(q)\ge\sum_i\mu_i$.
2107: Since $H\ge \bar H-\sum_i\mu_i$, $H(q)\ge0$.
2108: We conclude that $H(q)\ge0$ for all $\nu\in S$.
2109: In particular, $H(q)\ge0$ where $q$ is any
2110: superdifferential of $V$ at any interior point.
2111:
2112: Finally, consider a point $x\in G\cap\pl\R_+^J$. Any superdifferential
2113: of $V$ at $x$ is given as
2114: $q=\sum_{i\in I(x)}\eta_ie_i-\sum_{j=1}^J\nu_j\al_je_j$,
2115: where $\eta_i\ge0$.
2116: If $\max_{i\in I(x)}\lan q,\gamma_i\ran\ge0$, then
2117: (\ref{eq:subs}) holds. Otherwise, $\lan q,\gamma_i\ran<0$
2118: for all $i\in I(x)$. Consequently, any $q$ of the form above
2119: is given as $-\sum_{j=1}^J\nu_j'\al_je_j$, with $\nu'\in S$.
2120: As we have shown, in this case, $H(q)\ge0$. Therefore
2121: (\ref{eq:subs}) holds.
2122:
2123: Similarly, any subdifferential of $V$ at $x\in G\cap\pl\R_+^J$ is
2124: of the form $-\sum_{i\in I(x)}\eta_ie_i-\sum_{j=1}^J\nu_j\al_je_j$.
2125: In particular, $\lan q,e_i\ran\le0$ for all $i$, and (\ref{eq:supers})
2126: holds.
2127: \qed
2128:
2129: \section{Proofs of lemmas}\label{sec:lemmas}
2130:
2131: \noi{\bf Proof of Lemma \ref{lem:dpe}:}
2132: Let
2133: \[
2134: W^{n}(x)\doteq \inf E_{x}^{u,n}e^{-nc\sigma _{n}}.
2135: \]
2136: Since $c>0$, $W^{n}$ is well defined.
2137: Standard iterative methods can be used to construct a solution to the DPE
2138: \begin{equation}\label{eq:dpe3}
2139: 0=\inf_{u\in U}\left[ \til {\cal L}^{n,u}\bar W^{n}(x)-nc\bar W^{n}(x)\right] ,x\in G^{n}
2140: \end{equation}
2141: and the boundary condition $\bar W^{n}(x)=1$ if $x\notin G^{n}$.
2142: We claim that this solution coincides with the risk-sensitive cost.
2143: To see this,
2144: consider a controlled Markov process $(X^n,u)$ that starts at $x$.
2145: Then
2146: \[
2147: Y(t) \doteq \bar W^n(X^n(t))-\bar W^n(x)
2148: -\int_0^t\tilde{\cal L}^{n,u(s)}\bar W^{n}(X^n(s))ds
2149: \]
2150: is a martingale.
2151: Equation (\ref{eq:dpe3}) implies
2152: ${\cal L}^{n,u(s)}\bar W^{n}(X^n(s))\geq nc\bar W^{n}(X^n(s))$,
2153: and so
2154: \[
2155: \bar W^n(X^n(t))-\bar W^n(x)-\int_0^t nc \bar W^{n}(X^n(s))ds
2156: =\int_0^tZ(s)ds+Y(t)
2157: \]
2158: for some nonnegative process $Z$.
2159: Using Gronwall's lemma we obtain that for each $t<\infty$
2160: \[
2161: E^{n,u}_x \bar W^n(X^n(t\wedge \sigma^n))e^{-nc(t\wedge \sigma^n)}\geq \bar W^n(x),
2162: \]
2163: and by the Lebesgue Dominated Convergence Theorem
2164: \[
2165: E^{n,u}_x e^{-nc\sigma^n}\geq \bar W^n(x).
2166: \]
2167: If we define $u$ in terms of the feedback control that minimizes in (\ref{eq:dpe3}) then all the inequalities above become equalities, thus showing that $\bar W^n=W^n$.
2168:
2169:
2170:
2171: The definition of $W^{n\,}$ implies $W^{n}(x)=\exp \left[ -nV^{n}(x)\right] $
2172: . \ If we insert this into the DPE of $\bar W^{n}$ and multiply by $\exp \left[
2173: nV^{n}(x)\right] $ then the equation
2174: \begin{eqnarray*}
2175: 0& =& \inf_{u\in U}\left[ \sum_{j=1}^{J}n\lambda _{j}\left( \exp \left[
2176: -nV^{n}\left( x+\frac{1}{n}v_{j}\right) +nV^{n}\left( x\right) \right]
2177: -1\right) \right. \\ && \left. \mbox{} +\sum_{i=1}^{J}n\mu _{i}u_{i}\left( \exp \left[ -nV^{n}\left( x+
2178: \frac{1}{n}\pi (x,\tilde{v}_{i})\right) +nV^{n}\left( x\right) \right]
2179: -1\right) - nc\right]
2180: \end{eqnarray*}
2181: results. \ Recall the definition $l(x)=x\log x-x+1$ for $x>0$. We now
2182: divide throughout by $n$ and
2183: use the convex duality relation
2184: \[
2185: \left[ e^{y}-1\right] =\sup_{x>0}[xy-l(x)]
2186: \]
2187: to represent the terms in the previous display. \ For example, in the sum on
2188: $j$ we take $x=\bar{\lambda}_{j}/\lambda _{j}$ and $y=-\left[
2189: nV^{n}\left( x+\frac{1}{n}v_{j}\right) -nV^{n}\left( x\right) \right] $.
2190: Representing each term in this way and multiplying by $-1$ produces the
2191: first line in (\ref{eq:dpe}). \ The boundary condition that is the second line in (\ref{eq:dpe})
2192: follows directly from the relation between $W^{n}$ and $V^{n}$.
2193: \qed
2194:
2195:
2196:
2197: \noi{\bf Proof of Lemma \ref{lem:unif}:}
2198: We reduce the Lipschitz property on $(n^{-1}\Z^J_+)\cap\bar G$
2199: to a Lipschitz property near the boundary.
2200: To this end we use the following coupling.
2201: For $z\in G^n$, let $u^n(z)$ be a minimizer
2202: in (\ref{eq:dpe3}). Given a point $x$ on the lattice, let $X^x$ denote
2203: the process corresponding to the generator $\calL^{n,u^n}$ and starting at $x$
2204: (see the discussion following (\ref{eq:gens})).
2205: To simplify the notation we will not explicitly denote the dependence of quantities such as $X^x$ on $n$.
2206: Let $u(t)=u^n(X^x(t))$, and let $F_t$ be the filtration generated by $X^x$.
2207:
2208: Fix a point $ y \not= x $ and let $ X^y $ denote the queueing process
2209: on this probability space that
2210: starts at $ y $ and uses the control $u$.
2211: In other words,
2212: $X^y$ is the image,
2213: under the Skorokhod map,
2214: of $y+X^x(\cdot)-x$.
2215: The evolution of the processes $X^x$ and $X^y$ are identical, save that jumps which would cause $X^y$ to leave $\Z^J_+$ are deleted.
2216: Automatically, $u$ is suboptimal for the control problem starting from $y$.
2217: Define
2218: \be
2219: V^n(y ; \policy ) = - n^{-1}\log E_x^{\policy ,n} e^{-nc\sig^y }
2220: \ee
2221: where $ \sig^y $ is the exit time of $ X^y $ from $G$.
2222: Note that due to the coupling we may take expectations with respect
2223: to $ E_x^{u,n} $ rather then with respect to $ E_y^{u,n} $.
2224: Since $(X^y,u)$ is a (possibly suboptimal) controlled Markov process, we have
2225: \be\label{eq:VbyVu}
2226: V^n (x ) - V^n ( y ) \leq V^n (x) - V^n (y ; \policy ) .
2227: \ee
2228:
2229:
2230: Define $ \sig = \min \{ \sig^x , \sig^y \} $.
2231: By Theorem~\ref{th:SP} on the Lipschitz continuity
2232: of the Skorokhod map we have
2233: \be
2234: \dist ( X^x (\sig ) , \partial G ) \leq K_1 | x - y | , \quad
2235: \dist ( X^y (\sig ) , \partial G ) \leq K_1 | x - y |,
2236: \ee
2237: since at least one of the processes has left $G$ by $ \sig $.
2238: In the last display, $K_1$ is the constant appearing in (\ref{sp:lip}). We claim that
2239: \be\label{dist:1}
2240: V^n (x ) - V^n ( y ) \leq \sup \{ V^n (z) : z \in S \} \ee
2241: where $ S \doteq \{z\in n^{-1}\Z_+^J\cap \bar G:
2242: \dist (z , \pl_{co} G ) \leq K_1 |x-y| \} $.
2243: To establish this, note that
2244: \beaa
2245: V^n (x ) - V^n ( y , \policy )
2246: &=& - \frac 1 n \left [ \log E_x^{u,n} e^{-nc \sigma^x }
2247: - \log E^{u,n}_x e^{-nc \sigma^y } \right ] \\
2248: & \leq & - \frac 1 n \left [ \log E_x^{u,n} \left [ e^{-nc \sigma }
2249: E_x^{u,n} \left ( e^{-nc (\sigma^x - \sigma )} \left | X^x (\sigma)
2250: \right . \right ) \right ] - \log E_x^{u,n} e^{-nc \sigma} \right ] \\
2251: & \leq & \sup_{z\in S} - \frac 1 n \left [ \log E_x^{u,n} \left [ e^{-nc \sigma }
2252: E_x^{u,n} \left ( e^{-nc (\sigma^x - \sigma )}
2253: \left | X^x (\sigma) = z \right . \right ) \right ]
2254: - \log E_x^{u,n} e^{-nc \sigma} \right ]\\
2255: & = & \sup_{z\in S} - \frac 1 n \left [ \log \left [ E_x^{u,n}
2256: \left ( e^{-nc (\sigma^x - \sigma )} \left | X^x (\sigma) = z
2257: \right . \right ) E_x^{u,n} e^{-nc \sigma } \right ]
2258: - \log E_x^{u,n} e^{-nc \sigma} \right ]\\
2259: & = & \sup_{z\in S} - \frac 1 n \left [ \log E_x^{u,n} \left (
2260: e^{-nc (\sigma^x - \sigma )} \left | X^x (\sigma) = z
2261: \right . \right ) \right ]
2262: \eeaa
2263: However, by the strong Markov property,
2264: $$
2265: - \frac 1 n \left [ \log E_x^{u,n} \left ( e^{-nc (\sigma^x - \sigma )}
2266: \left | X^x (\sigma) = z \right . \right ) \right ]
2267: \leq \sup_u - \frac 1 n \log \left ( E_z^{u,n} e^{-nc \sigma^z }\right )
2268: = V^n (z)
2269: $$
2270: and together with~(\ref{eq:VbyVu}) we have~(\ref{dist:1}).
2271:
2272: To prove the lemma, one needs to show that
2273: $|V_n(x)-V_n(y)|\le c_0|x-y|$ for all $n$ and all
2274: $x,y\in(n^{-1}\Z_+^J)\cap\bar G$, where $c_0$ does not depend
2275: on $x,y$ and $n$. It suffices to prove this inequality for
2276: $x,y$ such that $|x-y|=n^{-1}$.
2277: Since the roles of $x $ and $y$ are symmetric, and in view of
2278: (\ref{dist:1}), it suffices to show that
2279: for $\{x\in G:\dist(x,\pl_{co} G)\le K_1n^{-1}\}$,
2280: \begin{equation}\label{eq:55}
2281: V_n(x)=-n^{-1}\log\inf_uE_x^{u,n}e^{-nc\sig^x}\le c_1n^{-1},
2282: \ee
2283: where $c_1 > 0$ is a constant.
2284:
2285: Let us first treat the case where $G$ is not a rectangle.
2286: In that case, Condition \ref{cond:G} implies that for
2287: any $x$ with $\dist(x,\pl_{co} G)\le K_1n^{-1}$,
2288: \be\label{imp:1}
2289: \mbox{ there is } i\in\calJ_+\ \mbox{ such that }\ x+c'n^{-1}e_i\not\in G,
2290: \ee
2291: where $c'$ is a constant.
2292: Let such $i$ be fixed.
2293: To show (\ref{eq:55}), it is enough to show that for any
2294: $x$ such that $\dist(x,\pl_{co}G)\le K_1n^{-1}$, and any
2295: $n$ and $u$,
2296: \begin{equation}\label{eq:56}
2297: E_x^{u,n}e^{-nc\sig^x}\ge c_2>0.
2298: \ee
2299: Recall that $\la_i>0$. Let $S_t$ denote the event that
2300: all service processes and all arrival processes, except for the
2301: one corresponding to $i$, do not increase on $[0,t]$.
2302: Recall that the expected time till a Poisson process of rate $\la$
2303: hits level $K$ is $K/\la$.
2304: Then for any $\al\in(0,1)$
2305: \beaa
2306: E_x^{u,n}e^{-nc\sig^x} & \ge & \al P_x^{u,n}(e^{-nc\sig^x}>\al)\\
2307: &=&
2308: \al P_x^{u,n}\lt(\sig^x<-\frac{\log\al}{nc}\rt).
2309: \eeaa
2310: Choosing $t_0=-(\log\al)/nc = 2\tilde c/n\la_i$
2311: and using $P_x^{u,n}(\sig^x<2E\sig^x)\ge1/2$,
2312: \beaa
2313: E_x^{u,n}e^{-nc\sig^x} & \ge & \al P_x^{u,n}(\sig^x<t_0|S_{t_0})P_x^{u,n}(S_{t_0})\\
2314: &\ge&
2315: e^{-2 c \tilde c / \la_i } \frac 1 2 c_3
2316: \eeaa
2317: where $c_3 > 0 $ is the probability that a Poisson process with rate
2318: $ n c_4 $ has not jumped by time $ t_0 = 2 \tilde c / \la_i n $.
2319: This proves (\ref{eq:56}), which implies (\ref{eq:55}),
2320: and hence the statement of the lemma holds.
2321:
2322: In the case where $G$ is a rectangle,
2323: the bound~(\ref{dist:1}) does not suffice since $V^n (x) $ is
2324: discontinuous near $\pl_cG$.
2325: We therefore prove that a similar bound applies, where there
2326: supremum is over $S=\{z\in(n^{-1}\Z_+^J)\cap\bar G:
2327: \dist(z,\pl_oG)\le K_1|x-y|\}$.
2328: To apply the previous argument we need to show that if $ X^x (t)$ is
2329: close to $\pl_cG$,
2330: then neither $ X^x $ nor $ X^y $ will exit (locally) through that boundary.
2331: This is clear for $ X^x $: the only
2332: way for the process to leave $G$ is due to a service to one of the queues,
2333: say, queue $j$, leading to an increase in queue $i$.
2334: However, allowing this service is certainly not optimal: it is better to
2335: avoid this control, as our objective is to increase $ \sigma^x $.
2336: To prevent $ X^y $ from exiting we need to modify the coupling argument as
2337: follows. The control $u^y$ used by $ X^y$ avoids a jump that leads
2338: $ X^y $ to exit through $\pl_cG$ (that is, queue $j$ above will not be
2339: served if $ X_i^y (t) = z_i - 1 $.) Note that this is the only possible
2340: type of jump that leads the process out of $G$.
2341: Moreover, the $ \ell_1 $ distance
2342: between $ X^x $ and $ X^y $ may only decrease due to this change in
2343: control: the control is changed only if $ X^x_i (t) < X^y_i (t) $,
2344: and following the service $ X^x_i $ increases by $1$ so that
2345: $ | X^x_i (t) - X^y_i (t) | $ decreases by $1$, while
2346: $ X^x_j $ decreases by $1$.
2347:
2348: Condition \ref{cond:G} still
2349: implies (\ref{imp:1}), but only for $x$ such that
2350: $$
2351: \mbox{$z_i-x_i\le n^{-1}$, for some $i\in\calJ_+$.}
2352: $$
2353: For such $x$, the argument in the last paragraph holds.
2354: However, for $x$ near $\pl_c G$
2355: there is nothing to prove, since the process never exits through such a
2356: boundary.
2357: \qed
2358:
2359:
2360: \noi{\bf Proof of Lemma \ref{lem:bound:m}:}
2361: The first part is an immediate consequence of the fact that
2362: $ u_i \geq 0 $ and both
2363: $\calL^{n,u,m}V^n(x)$ and $\rho(u,m)$ depend on $u$ as $\sum_iu_i\eta_i$,
2364: where $\eta_i$ is a function of $m_i ,x,n$ but not of $u$.
2365:
2366: For the second part of the lemma, one can explicitly solve for
2367: $m^n$ in terms of $V^n$, and get $m^n(x,u)=m^n(x)=
2368: ((\bar\la^n_i(x)),(\bar\mu^n_i(x)))$,
2369: where
2370: $$
2371: \bar\la^n_i(x)=\la_ie^{-n\del_i V^n(x)}, \quad
2372: \bar\mu^n_i=\mu_ie^{-n\tilde\del_i V^n(x)},
2373: $$
2374: and
2375: $$
2376: \del_i V^n(x)\doteq V^n(x+n^{-1}v_i)-V^n(x), \quad
2377: \tilde\del_i V^n(x)\doteq V^n(x+n^{-1}\pi(x,\tilde v_i))-V^n(x).
2378: $$
2379: The result follows from Lemma \ref{lem:unif},
2380: since it shows that there is a constant $b_2$
2381: independent of $x,n$ where
2382: $$
2383: n\del_i V^n(x) \ge-b_2, \quad
2384: n\tilde\del_i V^n(x) \ge-b_2.
2385: $$
2386: \qed
2387:
2388: \noi{\bf Proof of Lemma \ref{lem:cont}:}
2389: We fix $b\in [b^*,\infty]$ and suppress it from the notation throughout
2390: the proof.
2391: Item 1 of the lemma is trivial under Condition \ref{cond:G}.1.
2392: Under Condition \ref{cond:G}.2, by continuity of the functions
2393: $ \phi_i$, we only need to show is that $\pl_{co}G_a$ and $\pl_{co}G$
2394: do not intersect. Consider first $a>0$, and let $x\in\pl_{co}G_a$.
2395: Then $x_i=a+ \phi_i(x_1,\ldots,x_{i-1},x_{i+1},\ldots,x_J)$
2396: for some $i\in\calJ_+$, and therefore
2397: $x$ cannot belong to the closure of $G$. The proof for $a<0$ is
2398: similar.
2399:
2400: Let $\beta_0[u](t)=m_0$ for all $u,t$, where $m_0$ sets all $\bar\la_i=
2401: \la_i$ and $\bar\mu_i=0$. Then $\rho(u(t),\beta_0[u](t))$ is bounded
2402: by a constant, and the dynamics, unaffected by $u$, follow $ X(t)=
2403: x+\sum_i\la_ie_it$ and leave the bounded set $G$ within a finite time bounded
2404: by ${\rm diam}(G)/\max_i\la_i<\infty$. Therefore
2405: \[
2406: V^- (x)\le \sup_u C (x,u,m_0) = C(x,\beta_0) \le c_1< + \infty .
2407: \]
2408: Similarly,
2409: \[
2410: V^+ (x)\le \sup_\alpha C (x,\alpha (m_0),m_0) \le c_1 < + \infty .
2411: \]
2412:
2413: It is useful to notice that for all $a\in(0,a_0)$
2414: and $y\in\pl_o G$ there is $i=i_y\in\calJ_+$ such that
2415: $y+2ae_{i_y}\not\in G_a$. Similarly, for all $a\in(-a_0,0)$
2416: and $y\in\pl_o G_a$ there is $i=i_y\in\calJ_+$ such that
2417: $y+2ae_{i_y}\not\in G$.
2418:
2419: First consider $a>0$ and recall that $ \sigma_a $ (resp., $\sigma $)
2420: is the exit time from $G_a $ (resp., $G$), so that for any fixed $ u$ and
2421: $ \beta [u] $ we have $ \sigma \leq \sigma_a $.
2422: Therefore, since $c$ and $ \rho $ are positive,
2423: \beaa
2424: V^-_a(x)
2425: &= & \inf_\beta\sup_u \int_0^{\sig_a} (c+\rho(u(s), \beta[u](s))ds \\
2426: &\geq & \inf_\beta\sup_u \int_0^{\sig} (c+\rho(u(s), \beta[u](s))ds \\
2427: & = & V^- (x) .
2428: \eeaa
2429: Thus to prove the Lipschitz property a one-sided bound suffices.
2430: Recall that $ \phi (\sigma ) $ is the exit point from $ G$ and
2431: for each $\beta$ define the extension $\beta_a$ by
2432: $$
2433: \beta_a[u]=\lt\{\begin{array}{ll}\beta[u](t) & t\in[0,\sig), \\
2434: \hat m & t\in[\sig,\infty),
2435: \end{array}\rt.
2436: $$
2437: where $\hat m$ sets all $\bar\mu_j=0$ and
2438: $\bar\la_j=1_{j=i_{\phi(\sig)}}$.
2439: Then for any $\beta$
2440: \beaa
2441: V^-_a(x)
2442: &=& \inf_{\beta}\sup_u C_a(x,\beta[u],u) \\
2443: &\le& \inf_{\beta}\sup_u C_a(x,\beta_a[u],u)\\
2444: &=& \inf_\beta\sup_u\lt[C(x,\beta[u],u)+\int_\sig^{\sig_a}
2445: (c+\rho(u(s),\hat m)ds\rt] \\
2446: &\le&
2447: V^-(x)+c_1a,
2448: \eeaa
2449: where the last line follows since $\rho(u(s),\hat m)$ is bounded
2450: and since by the previous paragraph, $\sig_a-\sig\le 2a$.
2451: Note that $c_1$ does not depend on $b\in[b^*,\infty]$.
2452: For $a<0$ the same argument shows that $V^-(x)\le V^-_a(x)+c_3|a|$,
2453: by interchanging the roles of $G$ and $G_a$.
2454:
2455:
2456: For $ V_a^+ (x) $ note that an argument as above gives
2457: $ V_a^+ (x) \geq V^+ (x) $. For each $m$ define $ m_a $ by
2458: $$
2459: m_a = \lt\{ \begin{array}{ll} m (t) & t\in[0,\sig), \\
2460: \hat m & t\in[\sig,\infty),
2461: \end{array}\rt .
2462: $$
2463: where $ \hat m $ is as above. Let $ \alpha_\epsilon $ be an $\epsilon
2464: $-optimal strategy. Then, since for any fixed $ u$ and
2465: $ m $ we have $ \sigma \leq \sigma_a $,
2466: \beaa
2467: V^+_a(x)
2468: &=& \sup_\alpha \inf_m C_a (x,m, \alpha[m]) \\
2469: & \leq & \inf_m C_a (x,m, \alpha_\epsilon [m]) + \epsilon \\
2470: & \leq & \inf_m C_a (x,m_a , \alpha_\epsilon [m_a ]) + \epsilon,
2471: \eeaa
2472: since we are taking the infimum over a smaller class of controls.
2473: By the definition of $ C_a $,
2474: \beaa
2475: V^+_a(x)
2476: & \leq & \inf_m C (x, m , \alpha_\epsilon [m ])
2477: + \int_\sigma^{\sigma_a}
2478: (c+\rho(\alpha_\epsilon [m_a ](s),\hat m (s))\, ds + \epsilon \\
2479: & \leq & \sup_\alpha \inf_m C (x,m, \alpha[m]) + c_2 a + \epsilon
2480: \eeaa
2481: by the previous argument,
2482: where $c_2 $ does not depend on $x$, $ \epsilon $ and $b$.
2483: Since $ \epsilon $ is
2484: arbitrarily small, the proof for $a>0 $ and $ V_a^+ (x) $ is established.
2485: \qed
2486:
2487: \noi{\bf Proof of Lemma \ref{lem:apriori}:}
2488: We suppress $b$ from the notation, throughout the proof.
2489: It is obvious that one can restrict the infimum
2490: over $\beta\in B$ to the class of strategies
2491: $\beta$ for which $C(x,\beta)\le V^-(x)+1$.
2492: Within this class, for every $\beta$ and $u$,
2493: $$
2494: \sigma c \leq \int_0^\sig[c+\rho(u(t),\beta[u](t))]dt\le V^-(x)+1,
2495: $$
2496: and therefore one always has that $\sig\le T_0\doteq(V(x)+1)/c$.
2497: Lemma \ref{lem:apriori} asserts an upper bound on the cost till time a fixed time $T_0$,
2498: and so we must define the strategy for times $t \in [\sigma, T_0]$.
2499: Let $\hat m$ be an arbitrary fixed element of $M$.
2500: Then the extended $\beta$ is just
2501: $$
2502: \hat\beta[u](t)=\lt\{\begin{array}{ll}\beta[u](t) & t<\sig.\\
2503: \hat m & t\ge\sig,
2504: \end{array}\rt.
2505: $$
2506: With this definition
2507: one has that $C(x,u,\beta[u])=C(x,u,\hat\beta[u])$. One can
2508: therefore further restrict to strategies $\beta$ satisfying
2509: $\hat\beta=\beta$. For such $\beta$, it follows that
2510: $$
2511: \int_0^{T_0}\rho(u(t),\beta[u](t))dt \le c_1T_0,
2512: $$
2513: where $c_1$ does not depend on $u,\beta$.
2514: The result regarding $V^-$ follows.
2515:
2516: Regarding $V^+$,
2517: let $m_0$ be a control which sets all $\mu_i$ and $\la_i$ to
2518: zero, except that $\la_{i_0}=1$ for some $i_0\in\calJ_+$.
2519: Then for any $\al\in A$ and $m$ for which
2520: \begin{equation}\label{eq:20}
2521: C(x,\al[m],m)\le
2522: C(x,\al[m_0],m_0),
2523: \end{equation}
2524: one has $c\sig(x,\al[m],m)
2525: \le C(x,\al[m],m)\le
2526: C(x,\al[m_0],m_0)\le c_1<\infty$. Note that $c_1$ can be chosen
2527: independent of $\al$, since the dynamics and running cost
2528: under $m_0$ are independent of $\al$. Clearly, for each $\al$ it suffices
2529: to consider, in optimizing over $m$, only those $m$ that satisfy
2530: (\ref{eq:20}). It follows that it suffices to consider only
2531: those $m$ for which $\s(x,\al[m],m)\le c_1/c$.
2532: This completes the proof of the lemma.
2533: \qed
2534:
2535:
2536: \noi{\bf Proof of Lemma \ref{lem:vallip}:}
2537: Fix $ b\in[b^*,\infty]$ which we omit from the notation.
2538: Recall from Lemma \ref{lem:cont} that $V^\pm$ are bounded
2539: on $G$.
2540: We first show that $V^-$ is Lipschitz. Assume first that
2541: Condition \ref{cond:G}.2 holds.
2542: Recall that for $x\in G$,
2543: $$
2544: V^-(x)=\inf_\beta\sup_u C(x,\beta[u],u).
2545: $$
2546: Let $\beta_\eps^x$ be an $\eps$-optimal strategy starting
2547: from $x$, i.e.,
2548: $$
2549: \sup_u C(x,\beta_\eps^x[u],u)\le V^-(x)+\eps.
2550: $$
2551: For any $z\in G$ let $\s_z=\inf\{t:\phi_z\not\in
2552: G\}$, where $\phi_z$ is the solution to
2553: $\dot\phi=\pi(\phi,v(u,\beta_\eps^x[u]))$, with $\phi(0)=z$.
2554: Note that $C(x,u,\beta_\eps^x[u])=\int_0^{\s_x}[c+\rho(u(t),
2555: \beta_\eps^x[u](t))]dt$ (with possibly $ \sigma_x = \infty $).
2556: Now let $y\in G$. Note that on $[0,\s_x\w\s_y]$, one has by
2557: the Lipschitz property of the Skorokhod map
2558: that $|\phi_x (t) -\phi_y (t)|\le c_1|x-y|$,
2559: where $c_1$ is some constant.
2560: Recall that we are considering the case of Condition \ref{cond:G}.2.
2561: Therefore, at $\s_x\w \s_y$, both
2562: $\phi_x$ and $\phi_y$ are within a distance of $c_1|x-y|$ of
2563: the boundary $\pl_o G$. Because of the assumptions on the domain $G$,
2564: there exists a constant $c_2$
2565: such that at time $\s_x\w\s_y$, both
2566: $\phi_x+c_2e_{i^*}\not\in G$ and
2567: $\phi_y+c_2e_{i^*}\not\in G$, where $i^*\in\calJ_+$
2568: and, moreover, $c_2 $ is
2569: independent of $ x,y, b \geq b^* $ and $ i^* $.
2570:
2571: Define $\beta_\eps^{x,y}$ as
2572: $\beta_\eps^{x,y}[u]=\beta_\eps^x[u]$ on $[0,\sig_x)$, and,
2573: if $\sig_y\ge\sig_x$, set $\beta_\eps^{x,y}[u]=m_0$ on $[\s_x,\s_y]$.
2574: Here, $m_0$ sets all $\bar\mu_i$ and all $\bar\la_i$ to zero, except
2575: that it sets $\bar\la_{i^*}=\la_{i^*}$, where $i^*$ is as above.
2576: Consequently, $ \sig_y < \sig_x + c^\prime | x - y | $ for some
2577: $ c^\prime > 0 $, and there exists $u_\eps$ such that
2578: \beaa
2579: V^-(y) &\le& \sup_u C(y,u,\beta_\eps^{x,y}[u]) \\
2580: &\le&
2581: \int_0^{\s_y}[c+\rho(u_\eps(s),\beta_\eps^{x,y}[u_\eps](s))]ds+\eps\\
2582: &\le&
2583: \int_0^{\s_x}[c+\rho(u_\eps(s),\beta_\eps^{x,y}[u_\eps](s))]ds
2584: +1_{\s_x<\s_y}
2585: \int_{\s_x}^{\s_y}[c+\rho(u_\eps(s),\beta_\eps^{x,y}[u_\eps](s))]ds
2586: +\eps\\
2587: &\le&
2588: C(x,u_\eps,\beta_\eps^{x}[u_\eps])+c_3|x-y|+\eps\\
2589: &\le&
2590: \sup_u C(x,u,\beta_\eps^x[u])+c_3|x-y|+\eps\\
2591: &\le&
2592: V^-(x)+c_3|x-y|+2\eps.
2593: \eeaa
2594: Since $x,y\in G$ and $\eps>0$ are arbitrary, and $c_3$ does not
2595: depend on them or on $ b$, $V^-$ is Lipschitz, uniformly for
2596: $ b\in[b^*,\infty] $.
2597:
2598: In case that Condition \ref{cond:G}.1 holds, the same
2599: argument shows that $V^-_a(y)\le V^-(x)+c_3|x-y|+2\eps$,
2600: where $a=c_1|x-y|$. By Lemma \ref{lem:cont}, this implies that
2601: $V^-(y)\le V^-(x)+c_4|x-y|+2\eps$, some constant $c_4$, and therefore
2602: $V^-$ is Lipschitz.
2603:
2604: Next, consider the upper value
2605: $$
2606: V^+(x)=\sup_\al\inf_mC(y,\al[m],m)
2607: $$
2608: under Condition \ref{cond:G}.2.
2609: Let $x,y\in G$. Note that there is an $\al_\eps^x$ such that
2610: $$
2611: V^+(x)\le\inf_m C(x,\al_\eps^x[m],m)+\eps.
2612: $$
2613: and an $m_\eps=m_\eps(x,y)$ for which
2614: \beaa
2615: V^+(y) &\ge& \inf_m C(y,\al_\eps^x[m],m) \\
2616: &\ge& C(y,\al_\eps^x[m_\eps],m_\eps)-\eps.
2617: \eeaa
2618: Let $\s_z=\inf\{t:\phi_z\not\in G\}$, where $\phi_z$ is the solution
2619: to $\dot\phi=\pi(\phi,v(\al_\eps^x[m_\eps],m_\eps))$, with
2620: $\phi(0)=z$. Let $i^*$ be defined in an analogous way to that in
2621: the first paragraph of the proof. Now
2622: define $\bar m_\eps=\bar m_\eps(x,y)$ as follows. If
2623: $\sig_x\le\sig_y$, let $\bar m_\eps=m_\eps$. If $\sig_y<\sig_x$, let
2624: $\bar m_\eps$ agree with $m_\eps$ on $[0,\sig_y)$ and with
2625: $m_0$ on $[\sig_y,\s_x]$. Here, $m_0$ sets all $\bar\mu_i$
2626: and all $\bar\la_i$ to zero, except that it sets $\bar\la_{i^*}
2627: =\la_{i^*}$.
2628: Since $m_\eps$ and $\bar m_\eps$ agree on $[0,\s_y)$, the restrictions
2629: to $[0,\sig_y]$ of $\al_\eps^x[m_\eps]$ and of $\al_\eps^x[\bar m_\eps]$
2630: agree a.e.\ on $[0,\sig_y]$, and therefore,
2631: $$
2632: C(y,\al_\eps^x[m_\eps],m_\eps)=C(y,
2633: \al_\eps^x[\bar m_\eps],\bar m_\eps).
2634: $$
2635: Arguing again by the Lipschitz property of the
2636: Skorokhod map and the definition of $m_0 $,
2637: there is a constant $c_4$ for which
2638: $(\sig_x-\sig_y)^+\le c_4|x-y|$. Hence
2639: \beaa
2640: V^+(y) &\ge& C(y,\al_\eps^x[\bar m_\eps],\bar m_\eps)-\eps \\
2641: &\ge & \int_0^{\sig_x}[c+\rho(\al_\eps^x[\bar m_\eps^x](s),\bar m_\eps^x(s))]
2642: ds - 1_{\sig_y<\sig_x}\int_{\sig_y}^{\sig_x}
2643: [c+\rho(\al_\eps^x[\bar m_\eps^x](s),\bar m_\eps^x(s))]ds -\eps\\
2644: &\ge& C(x,\al_\eps^x[\bar m_\eps],\bar m_\eps)-c_5|x-y|-\eps\\
2645: &\ge& \inf_m C(x,\al_\eps^x[m],m)-c_5|x-y|-\eps\\
2646: &\ge& V^+(x)-c_5|x-y|-2\eps.
2647: \eeaa
2648: Since $c_5$ does not depend on $x,y,\eps$ or $b$, we have that $V^+$
2649: is Lipschitz uniformly for $b\in[b^*,\infty] $.
2650:
2651: Under Condition \ref{cond:G}.1, the same argument shows that
2652: $V^+_a(y)\ge V^+(x)-c_5|x-y|-2\eps$, where $a=c_1|x-y|$, and
2653: again one argues by Lemma \ref{lem:cont}.
2654: \qed
2655:
2656:
2657: \noindent{\bf Proof of Lemma \ref{lem:xbar}:}
2658: The processes are constructed recursively using a sequence
2659: of standard exponential clocks.
2660: Recall that $\calL^{n,u,m}$ is given for every $n$, $u\in U$,
2661: $m\in M$ by
2662: \[
2663: \calL^{n,u,m}f(x) =
2664: \sum_{j=1}^J n \bar \lambda_j [f(x+{n}^{-1} v_j)-f(x)]
2665: + \sum_{i=1}^J n \bar \mu_i u_i [f(x+n^{-1}\pi(x,\tilde v_i))-f(x)].
2666: \]
2667: Given $n$, $x_n$ and $\beta$, we construct a filtered
2668: probability space and three processes, $\bar X(t)$, $\bar u(t)$
2669: and $\bar m(t)$ (to simplify notation, we do not write the superscript
2670: $n$ in the notation of $\bar X^n$, $\bar u^n$ and $\bar m^n$)
2671: such that (a) $\bar X,\bar u$ and $\bar m$
2672: are $(\bar F_t)$-adapted;
2673: (b) $\bar m(t)=\beta[\bar u](t)$ a.e.\ $t\ge0$, a.s.;
2674: (c) $\bar u(\cdot )=u^n(X(\cdot ))$ a.s.\ (where $u^n$ is as in
2675: the statement before the lemma); and (d) for any $f$, the process
2676: $$
2677: f(\bar X(t))-\int_0^t\calL^{n,\bar u(s),\bar m(s)}f(\bar X(s))ds
2678: $$
2679: is an $(\bar F_t)$-martingale.
2680: For (a--d) to hold, it suffices that (a--c) hold, and (e)
2681: on any finite interval the process $\bar X$ jumps finitely many
2682: times---we denote
2683: the $k$th jump by $\tau_k$ and let $\tau_0=0$; (f) the random times
2684: $(\tau_k)$ are
2685: stopping times on $(\bar F_t)$, and (g) denoting $X_k=\bar X(\tau_k)$,
2686: for any $k$,
2687: $$
2688: \bar E[f(X_{k+1})-f(X_k)|\bar F_{\tau_k}] =
2689: \sum_{i=1}^J \bar E[A_i^{k,\bar u,\bar m}
2690: +B_i^{k,\bar u,\bar m}|\bar F_{\tau_k}],
2691: $$
2692: where
2693: $$
2694: A_i^{k,\bar u,\bar m}= n\int_{\tau_k}^{\tau_{k+1}}
2695: \bar\la_i(s)ds [f(X_k+n^{-1}v_i)-f(X_k)],
2696: $$
2697: $$
2698: B_i^{k,\bar u,\bar m} = n\int_{\tau_k}^{\tau_{k+1}} \bar\mu_i(s)u_i(s)ds
2699: [f(X_k+n^{-1}\pi(X_k,\tilde v_i))-f(X_k)].
2700: $$
2701:
2702: The construction is recursive. On a complete probability
2703: space $(\bar\Om,\bar F,\bar P)$ we are given $2J$ independent
2704: i.i.d.\ standard Poisson processes, denoted
2705: $a_i$ and $b_i$, $i=1,\ldots,J$. Let $T^a_i(k)$ [resp., $T^b_i(k)$]
2706: denote the first time $a_i$ [resp., $b_i$] equals $k$.
2707: For each $\om\in\Om$ we construct recursively a sequence of times
2708: $(\tau_k)$ and the processes $\bar X, \bar u$ and $\bar m$
2709: up to time $\tau_k$. Once these processes are
2710: defined, we will define $(\bar F_t)$, $\bar F_t\subset \bar F$,
2711: $t\ge0$, and verify
2712: that items (a--c), (e--g) are satisfied on
2713: $(\bar\Om,\bar F,(\bar F_t),\bar P)$.
2714:
2715: We set $\bar X(0)=x_n$ and $\bar u(0)=u^n(x_n)$.
2716: Since $\bar m$ need only be defined
2717: almost everywhere on $[0,\infty)$, we do not define it at zero nor
2718: at any $\tau_k$, $k=1,2,\ldots$.
2719: Now assume that we have constructed $\tau_i$, $i\le k$ as well as
2720: the processes $\bar X$ and $\bar u$ on
2721: $[0, \tau_k]$ and $\bar m$ a.e.\ on $[0,\tau_k]$.
2722: Let
2723: $\hat u^k(t)=u^n(\bar X(t\w \tau_k))$, $t\ge0$. Let also
2724: $\hat m^k=\beta[\hat u^k]$.
2725: With $\hat u^k(\cdot )=(\hat u_i^k(\cdot ))$ and
2726: $\hat m^k(\cdot )=((\hat\la_i^k(\cdot )),(\hat\mu_i^k(\cdot )))$, let
2727: $$
2728: p_i^k(t)=n\int_0^t\hat\la_i^k(s)ds,
2729: \qquad q_i^k(t)=n\int_0^t\hat\mu_i^k(s)\hat u_i^k(s)ds,
2730: \qquad i=1,\ldots,J, \ t\ge0.
2731: $$
2732: Denoting $\Del z(s)=z(s)-z(s-)$, let also
2733: $$
2734: \tau_{k+1}=\inf\{t>\tau_k: \mbox{ either $\Del a_i(p_i^k(t))>0$
2735: or $\Del b_i(q_i^k(t))>0$ for some $i=1,\ldots,J$}\},
2736: $$
2737: where $\inf\emptyset=+\infty$.
2738: We first consider the case that $\tau_{k+1}<+\infty$.
2739: In this case,
2740: \be\label{eq:exi}
2741: \mbox{there is $i$ such that either $\Del a_i(p_i^k(\tau_{k+1}))>0$ or
2742: $\Del b_i(q_i^k(\tau_{k+1}))>0$.}
2743: \ee
2744: In the former case we let
2745: $\hat v^k= v_i$; otherwise we let $\hat v^k=\tilde v_i$.
2746:
2747: The three processes are defined on the next interval
2748: as follows. Let $\bar X(t)=\bar X(\tau_k)$ for $t\in(\tau_k,\tau_{k+1})$,
2749: and $\bar X(\tau_{k+1})=\bar X(\tau_k)+n^{-1}\pi(\bar X(\tau_k),\hat v^k)$.
2750: Let $\bar u(t)=u^n(\bar X(t))$ for $t\in(\tau_k,\tau_{k+1}]$.
2751: Let $\check u(t)=u^n(\bar X(t\w\tau_{k+1}))$ and define
2752: $\bar m(t)=\beta[\check u](t)$, $t\in[0,\tau_{k+1}]$.
2753: Note that since $\beta$ is a strategy,
2754: this definition of $\bar m$ is consistent with its definition
2755: up to $\tau_k$ since so is the definition of $\bar u$.
2756: For the same reason, for a.e.\ $t\le\tau_{k+1}$,
2757: $\bar m(t)=\hat m^k(t)$.
2758: In particular, the equations for $p_i^k, q_i^k$ still hold
2759: if we replace hats by bars, namely,
2760: \be\label{eq:pq}
2761: p_i^k(t)=n\int_0^t\bar\la_i(s)ds,
2762: \quad q_i^k(t)=n\int_0^t\bar\mu_i(s)\bar u_i(s)ds,
2763: \quad i=1,\ldots,J, \ \tau_k\le t\le \tau_{k+1}.
2764: \ee
2765: Note that the above relations are consistent in the sense that
2766: for a given $k$, they hold not only for
2767: $t\in[\tau_k,\tau_{k+1}]$, but in fact for $t\in[0,\tau_{k+1}]$.
2768: Hence, on the event $\tau_k\to\infty$, one can equivalently
2769: consider the processes
2770: \be\label{def:pq}
2771: p_i(t)=n\int_0^t\bar\la_i(s)ds,
2772: \quad q_i(t)=n\int_0^t\bar\mu_i(s)\bar u_i(s)ds,
2773: \quad i=1,\ldots,J, \ t\ge0.
2774: \ee
2775: This completes the definition of the three processes on $[0,\tau_{k+1}]$.
2776:
2777: In case that $\tau_{k+1}=+\infty$, the definitions above of
2778: $\bar X$, $\bar u$ and $\bar m$
2779: all apply on $(\tau_k,\tau_{k+1})$ and there is
2780: nothing else to define.
2781:
2782: To complete the construction of the three processes on
2783: $\bar\Om\times[0,+\infty)$, we must consider the set $\Om_0$
2784: of $\om\in\bar\Om$
2785: for which $\bar T\doteq\sup\tau_k$ is finite.
2786: We show that this set is $\bar P$-null
2787: owing to the fact that the range $\bar M^b $ of $\beta$ consists
2788: of bounded functions.
2789: Suppose $\bar T$ is finite. The construction above defines $\bar X,\bar u$
2790: and $\bar m$
2791: on $[0,\bar T)$. Let $\bar u'(t)=\bar u(t)$ for $t<\bar T$
2792: and define $\bar u'(t)$ arbitrarily
2793: on $[\bar T,+\infty)$ but such that $\bar u'\in\bar U$. Then $\bar m'
2794: =\beta[\bar u']$
2795: agrees with $\bar m$ a.e.\ on $[0,\bar T]$. Since each component of $\bar m'$
2796: is bounded by $b$,
2797: \be\label{eq:boundsum}
2798: n^{-1}\max_{i=1}^J [p_i(\bar T)\vee q_i(\bar T)]\le
2799: 2J\bar T b <+\infty.
2800: \ee
2801: However, by construction, $\bar T<\infty$ implies that
2802: either $a_i(p_i(t))\to\infty$ or
2803: $b_i(q_i(t))\to\infty$ as $t\uparrow\bar T$, for some $i$.
2804: Hence $\bar T<\infty$ must be a null set.
2805: We let $\bar X,\bar u$ and $\bar m$ be defined arbitrarily on $\Om_0$.
2806:
2807: The definition of the process $\bar Y$ is similar to that of
2808: $\bar X$, but where $\pi(x,v)$ is replaced by $v$ throughout.
2809: The relation $\bar X=\Gamma(\bar Y)$ is clear from the construction.
2810:
2811: Define for each $t\ge0$ $\bar F_t$ to be the $\sig$-field
2812: generated by $\{ \bar Y(s), s\in[0,t] \} $. Note that it is
2813: equivalently defined as the $\sig$-field generated by
2814: $\{a_i(p_i(t)), b_i(q_i(t)), i=1,\ldots,J\}$, where
2815: $p_i,q_i$ are as in (\ref{def:pq}).
2816: By construction, $\bar u(t)=u^n(\bar X(t))$,
2817: $t\ge0$ and item (c) holds. Item (b), namely that $\bar m=\beta[\bar u]$,
2818: also holds by construction.
2819: $\bar X$ and $\bar u$ are therefore $(\bar F_t)$-adapted,
2820: and since $\beta$ is a strategy, so is $\bar m$, and item (a) holds.
2821: Items (e) and (f) are trivial. Concerning (g),
2822: let $i^k\in\{1,\ldots,2J\}$ denote the index $i$ satisfying
2823: (\ref{eq:exi}) in case that $\Del a_i(p_i^k(t))>0$ holds,
2824: and let it denote $i+J$ in the case $\Del b_i^k(q_i^k(t))>0$.
2825: It suffices to show that for every $i\in\{1,\ldots,2J\}$,
2826: $$
2827: \bar P(i^k=i|\bar F_{\tau_k})=
2828: \lt\{\begin{array}{ll}
2829: \bar E[\int_{\tau_k}^{\tau_{k+1}}p_i(s)ds|\bar F_{\tau_k}]/Z_k
2830: & i\le J,\\[.1in]
2831: \bar E[\int_{\tau_k}^{\tau_{k+1}}q_{i-J}(s)ds|
2832: \bar F_{\tau_k}]/Z_k & i> J,
2833: \end{array}\rt.
2834: $$
2835: where $Z_k$ is a normalization factor (not depending on $i$).
2836: For $k=0$ ($\tau_k=0$),
2837: this is a well known property of exponential clocks.
2838: For $k>0$, the same argument holds, merely because conditional on
2839: $\bar F_{\tau_k}$, the processes $\int_{\tau_k}^\cdot p_i(s)ds$,
2840: $\int_{\tau_k}^\cdot q_i(s)ds$ are independent, and moreover,
2841: $a_i(\cdot-\tau_k)-a_i(\tau_k),b_i(\cdot-\tau_k)-b_i(\tau_k)$
2842: are still independent Poisson processes (which is a statement
2843: on the lack of memory for exponential random variables).
2844:
2845: The proof of the claim regarding the martingale associated with
2846: $\calL_0$ is similar (only simpler). This completes the proof of
2847: the first part of the lemma.
2848:
2849: Clearly,
2850: $$
2851: \max_i p_i(T_0)\vee q_i(T_0)\le nT_0b,
2852: $$
2853: where $T_0$ is as in Lemma \ref{lem:apriori}.
2854: Thus, if $N_n=\max\{k:\tau_k\le T_0\}$, then
2855: $$
2856: N_n\le \sum_i a_i(nT_0b)+b_i(nT_0b),
2857: $$
2858: and (\ref{ineq:Nn}) follows.
2859: \qed
2860:
2861: \noi{\bf Proof of Lemma \ref{lem:xbar2}:}
2862: The proof is completely
2863: analogous to that of Lemma \ref{lem:xbar}, and is therefore
2864: omitted.
2865: \qed
2866:
2867: \noi{\bf Proof of Theorem \ref{th:b}:}
2868: By Theorems \ref{th:unique} and \ref{th:solve},
2869: $V^{b,-}=V^{b,+}$ for all $b\in[b^*,\infty)$. As a result,
2870: Theorem \ref{th:limit}, implies that
2871: $V^n\to V^{b,-}$ for all $b\in[b^*,\infty)$, as $n\to \infty$.
2872: In particular,
2873: $V^{b,-}$ does not depend on $b\in[b^*,\infty)$.
2874: It remains to show that for all $x$, $V^{b,-}(x)\to V^-(x)$
2875: and $V^{b,+}(x)\to V^+(x)$ as $b\to\infty$.
2876:
2877: \noi\uu{\em Proof that $V^{b,-}\to V^-$.}
2878: It is immediate from the definitions that $V^-\le V^{b,-}$.
2879:
2880: Let $\beta\in B$, and let $\s=\s(x,u,\beta)$ be the exit
2881: time of $\phi$ from $G$ where $\dot\phi=\pi(\phi,v(u,\beta[u]))$,
2882: $\phi(0)=x$. Let $\bar\beta$ be defined by
2883: $$
2884: \bar\beta[u](t)=\lt\{\begin{array}{ll}\min\{b,\beta[u](t)\}
2885: & t\le\s, \\ \hat m & t>\s,
2886: \end{array}\rt.
2887: $$
2888: where $\hat m$ sets all $\bar\mu_j=0$ and $\bar\la_j=1_{j=i_{\phi(\s)}}$,
2889: and the minimum is componentwise. It is clear that
2890: $\bar\beta$ is a strategy. Let $\bar u$ be any extension
2891: of $u$ to $[0,\infty)$, and denote by $\bar\phi$ and $\bar\s$ the
2892: dynamics and exit time corresponding to $x,\bar\beta,\bar u$.
2893: Recall that by (\ref{eq:lal}) $b$ is greater than all $\lambda_i$ and $\mu_i$.
2894: Thus
2895: \beaa
2896: C(x,\bar u,\bar \beta[\bar u]) &=&
2897: \int_0^\s(c+\rho(u,\bar\beta[u])ds
2898: +1_{\bar\s>\s}\int_\s^{\bar \s}(c+\rho(\bar u,\hat m))ds\\
2899: &\le&
2900: C(x,u,\beta[u])+c_1(\bar\s-\s)^+.
2901: \eeaa
2902: Moreover, by the Lipschitz property of the Skorokhod map, and denoting
2903: $\beta[u]=(\bar\la_i,\bar\mu_i)$,
2904: $$
2905: (\bar\s-\s)^+
2906: \le c_2|\phi(\s)-\bar\phi(\s)|\le
2907: c_2\int_0^\s\sum_i[(\bar\la_i-b)^+ +(u_i\bar\mu_i-b)^+]ds
2908: $$
2909: Since it is enough to consider $\beta$ for which (for any $u$)
2910: $\la_i$ and $u_i\mu_i$ are uniformly integrable over $[0,\s]$,
2911: we have that $(\bar\s-\s)^+\le \del(b)$, where $\del(b)\to0$
2912: as $b\to\infty$. This shows that
2913: $\lim_{b\to\infty}V^{b,-}(x)\le V^-(x)$.
2914:
2915: \noi\uu{\em Proof that $V^{b,+}\to V^+$.}
2916: It is immediate that $V^+\le V^{b,+}$.
2917:
2918: To show that $V^+(x)\ge \lim_{b\to\infty}V^{b,+}(x)$
2919: it is enough to show that for $b\ge b^*$ and $a$ small,
2920: $V^+(x)\ge V^{b,+}_{-a}(x)$.
2921: For any $m\in\bar M$ let $m^b$ denote the pointwise and componentwise
2922: truncation of $m$ at level $b$. For any $\al\in A$, let $\al^b\in A$
2923: be defined by $\al^b[m]=\al[m^b]$.
2924: We will write $m\in\bar M(\al,a)$ if $m,\al,a$ satisfy
2925: $C_a(x,\al[m],m)\le V^+_a(x)+1$.
2926: In the expression for $V^+_a(x)$,
2927: $$
2928: \sup_\al\inf_m C_a(x,\al[m],m),
2929: $$
2930: it is enough to consider $\al\in A$ and $m\in\bar M(\al,a)$
2931: (including for $a=0$).
2932: For such $\al,m$, the functions
2933: $\bar\la_i,u_i\bar\mu_i$ are uniformly integrable over $[0,T]$.
2934: Let $\al\in A$ and $m\in\bar M(\al^b,0)$. Consider
2935: a truncation of $\bar\la_i$ and $\bar\mu_i$ at $b$.
2936: Denote by $\phi$ [resp., $\phi^b$] the dynamics that correspond to
2937: $(x,\al^b,m)$, [resp., $(x,\al^b,m^b)$].
2938: Then the effect of the truncation
2939: on $\phi$ is such that for all $a>0$ there is $b$ such
2940: that $\sup_{[0,T]}|\phi-\phi^b|\le a$ (by uniform integrability).
2941: In particular, $|\phi^b(\s)-\phi(\s)|\le a$.
2942: Hence, using
2943: the monotonicity of the running cost for large values of
2944: the rates, and that $\al^b[m^b]=\al^b[m]$,
2945: $$
2946: C_{-a}(x,\al^b[m^b],m^b)\le C(x,\al^b[m],m).
2947: $$
2948: We thus have
2949: $$
2950: C_{-a}(x,\al[m^b],m^b)\le C(x,\al^b[m],m).
2951: $$
2952: Since $m\in\bar M(\al^b,0)$ implies that $m^b\in\bar M(\al^b,-a)$,
2953: \beaa
2954: \inf_{m\in\bar M^b}C_{-a}(x,\al[m],m) &\le&
2955: \inf_{m:m^b\in\bar M(\al^b,-a)}
2956: C_{-a}(x,\al[m^b],m^b)
2957: \\ &\le& \inf_{m\in\bar M(\al^b,0)}C(x,\al^b[m],m)
2958: \\ &=&
2959: \inf_{m\in\bar M}C(x,\al^b[m],m).
2960: \eeaa
2961: Hence
2962: $$
2963: \sup_{\al\in A}\inf_{m\in\bar M^b}C_{-a}(x,\al[m],m)\le
2964: \sup_{\al\in A}\inf_{m\in\bar M}C(x,\al[m],m).
2965: $$
2966: Taking $a\to0$ by letting $b\to\infty$, we have from
2967: Lemma \ref{lem:cont} that $\lim_bV^{b,+}(x)\le V^+(x)$.
2968: \qed
2969:
2970:
2971: \vspace{\baselineskip}
2972: \noindent
2973: {\it 1991 Mathematics Subject Classification.}
2974: Primary 60F10, 60K25;
2975: Secondary 93E20, 60F17.
2976:
2977:
2978: \bibliographystyle{plain}
2979: \begin{thebibliography}{99}
2980:
2981: \bibitem{atadup2} R.~Atar and P.~Dupuis.
2982: A differential game with constrained dynamics.
2983: To appear,
2984: {\it Nonlinear Analysis: theory, methods and applications.}
2985:
2986: \bibitem{atadupshw}
2987: R. Atar, P. Dupuis and A. Shwartz,
2988: ``Explicit solutions to a network control problem in
2989: the large deviation regime'', 1--18, preprint
2990:
2991: \bibitem{BDM} J.S.~Baras, A.J.~Dorsey and A.M.~Makowski,
2992: ``Two competing queues with geometric service requirements
2993: and linear costs: the $ \mu$-c rule is often optimal,"
2994: {\em Adv.\ Appl.\ Prob.\ \bf 17} pp.~186--209, 1985.
2995:
2996: \bibitem{balday1}
2997: J. Ball, M. Day, T. Yu and P. Kachroo,
2998: \newblock Robust L2-gain control for nonlinear systems with projection
2999: dynamics and input constraints: an example from traffic control.
3000: \newblock {\em Automatica}, 35:429--444, 1999.
3001:
3002: \bibitem{balday2}
3003: J. Ball, M. Day, and P. Kachroo,
3004: \newblock Robust feedback control for a single server queueing system.
3005: \newblock {\em Mathematics of Control, Signals, and Systems}, 12:307--345, 1999.
3006:
3007: \bibitem{barcap}
3008: M. Bardi and I. Capuzzo-Dolcetta.
3009: {\it Optimal control and viscosity solutions
3010: of Hamilton-Jacobi-Bellman equations.}
3011: Birkhauser, Boston. 1997
3012:
3013: \bibitem{buddup}
3014: A.~Budhiraja and P.~Dupuis.
3015: \newblock Simple necessary and sufficient conditions for the stability of
3016: constrained processes.
3017: \newblock {\em SIAM J. Applied Math.}, 59:1686--1700, 1999.
3018:
3019: \bibitem{capLion} I.~Capuzzo-Dolcetta and P.-L.~Lions,
3020: Hamilton-Jacobi equations with state constraints,
3021: {\em Trans.\ AMS \bf 318} pp.~643--683, 1990.
3022:
3023: \bibitem{delmey} C.~Dellacherie and P.-A.~Meyer. {\em Probabilit\'es
3024: et potentiel/ Th\'eorie des martingales}, Hermann, Paris, 1980.
3025:
3026: \bibitem{dupell}
3027: P. Dupuis and R.~S. Ellis.
3028: \newblock {\it A Weak Convergence Approach to the Theory of Large
3029: Deviations}.
3030: \newblock John Wiley \& Sons, New York, 1997.
3031:
3032: \bibitem{dupkus}
3033: P. Dupuis and H. Kushner.
3034: Minimizing escape probabilities: a large deviations approach.
3035: {\it SIAM J. Control Optim. 27} (1989), no. 2, 432--445
3036:
3037: \bibitem{dupish1}
3038: P.~Dupuis and H.~Ishii,
3039: On Lipschitz continuity of the solution mapping to the Skorokhod
3040: Problem, with applications.
3041: {\em Stochastics \bf 35}, pp.~31--62, 1991.
3042:
3043: \bibitem{dupish2}
3044: P.~Dupuis and H.~Ishii,
3045: On oblique derivative problems for fully
3046: nonlinear second-order elliptic PDE's
3047: on domains with corners.
3048: {\it Hokkiado U. Math. J.} {\bf 20}, pp. 135--164, 1991.
3049:
3050:
3051: \bibitem{dupjampet}
3052: P.~Dupuis, M.R. James and I.R.~Petersen.
3053: Robust properties of risk--sensitive control.
3054: {\em Math. of Control, Signals and Systems
3055: \bf 13} pp.~318--332, 2000.
3056:
3057: \bibitem{dupmce}
3058: P. Dupuis and W. M. McEneaney.
3059: Risk-sensitive and robust escape criteria.
3060: {\it SIAM J. Control Optim. 35} (1997), no. 6, 2021--2049.
3061:
3062: \bibitem{dupnag}
3063: P. Dupuis and A. Nagurney.
3064: Dynamical systems and variational inequalities.
3065: {\it Ann. Oper. Res.} 44 (1993), no. 1-4, 9--42.
3066:
3067: \bibitem{dupram23}
3068: P.~Dupuis and K.~Ramanan,
3069: Convex duality and the Skorokhod Problem. I, II.
3070: {\em Probability Theory and Related Fields \bf 2},
3071: pp~153--195, 197--236, 1999.
3072:
3073: \bibitem{ellkal} R.J.~Elliott and N.J.~Kalton.
3074: {\em The existence of value in differential games,}
3075: Memoirs of the American Mathematical
3076: Society, No. 126. American Mathematical Society,
3077: Providence, R.I., iv+67 pp., 1972.
3078:
3079: \bibitem{flesou}
3080: W. H. Fleming and P. E. Souganidis.
3081: PDE-viscosity solution approach to some problems of large deviations.
3082: {\it Ann. Scuola Norm. Sup. Pisa Cl. Sci.} (4) 13 (1986),
3083: no. 2, 171--192.
3084:
3085: \bibitem{harrei}
3086: J. M. Harrison and M. I. Reiman.
3087: Reflected Brownian motion on an orthant.
3088: {\it Ann. Probab.} 9 (1981), no. 2, 302--308.
3089:
3090: \bibitem{Klimov} G.P.~Klimov.
3091: Time sharing service systems I.
3092: {\em Theory Prob.\ Appl.} {\bf 19} pp.~532--551, 1974.
3093:
3094: \bibitem{reiwil} M.I.~Reiman and R.J.~Williams.
3095: A boundary property of semimartingale reflecting Brownian motions.
3096: {\it Probab. Theory Related Fields} 77 (1988), no. 1, 87--97.
3097:
3098: \bibitem{roc} R.T.~Rockafellar, {\em Convex analysis.} Princeton, NJ,
3099: 1970.
3100:
3101: \bibitem{shwe} A.~Shwartz and A.~Weiss,
3102: {\em Large deviations for performance analysis.} Chapman and Hall, 1995.
3103:
3104: \bibitem{son} H.M.~Soner. Optimal control with state space constraints I.
3105: {\em SIAM J.\ Control Opt.} {\bf 24} pp.~552--561, 1986.
3106:
3107: \bibitem{W} J.~Walrand,
3108: ``A note on `optimal control of a queueing system with two heterogeneous
3109: servers',''
3110: {\em System Control Lett.\ \bf 4} pp.~131--134, 1984.
3111:
3112: \end{thebibliography}
3113:
3114: \end{document}
3115:
3116:
3117:
3118: